Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2018 Jun 1.
Published in final edited form as: Curr Opin Genet Dev. 2017 Feb 14;44:1–8. doi: 10.1016/j.gde.2017.01.008

Designer protein disaggregases to counter neurodegenerative disease

James Shorter 1,*
PMCID: PMC5447488  NIHMSID: NIHMS852284  PMID: 28208059

Abstract

Protein misfolding and aggregation unify several devastating neurodegenerative disorders, including Alzheimer’s disease, Parkinson’s disease, and amyotrophic lateral sclerosis. There are no effective therapeutics for these disorders and none that target the reversal of the aberrant protein misfolding and aggregation that cause disease. Here, I showcase important advances to define, engineer, and apply protein disaggregases to mitigate deleterious protein misfolding and counter neurodegeneration. I focus on two exogenous protein disaggregases, Hsp104 from yeast and gene 3 protein from bacteriophages, as well as endogenous human protein disaggregases, including: (a) Hsp110, Hsp70, Hsp40, and small heat-shock proteins; (b) HtrA1; and (c) NMNAT2 and Hsp90. I suggest that protein-disaggregase modalities can be channeled to treat numerous fatal and presently incurable neurodegenerative diseases.

Introduction

Deleterious protein misfolding and aggregation underpin several invariably fatal and age-related neurodegenerative diseases, including Alzheimer’s disease (AD), Parkinson’s disease (PD), and amyotrophic lateral sclerosis (ALS) [1]. Typically, in each disease specific proteins misfold, aggregate, and wreak havoc on the nervous system [1]. In AD, amyloid-β (Aβ) peptides form extracellular, neuritic plaques and tau forms intracellular neurofibrillary tangles in afflicted neurons [1]. By contrast, in PD, α-synuclein (α-syn) forms cytoplasmic Lewy bodies in degenerating dopaminergic neurons [1]. In most ALS cases, RNA-binding proteins with prion-like domains, such as TDP-43 or FUS, mislocalize from the nucleus to cytoplasmic aggregates in degenerating motor neurons and glia [1,2]. Current treatments for these disorders are palliative and ineffective. No therapeutics exist that reverse the aberrant protein misfolding and aggregation that underlie disease. The lack of effective therapies is a cause of immense angst as these diseases are increasing in prevalence as our population ages.

Complexity of protein misfolding

Protein misfolding is a complex, multistate process [3,4]. The specific proteins that misfold in neurodegenerative disease are often intrinsically disordered, harbor an intrinsically disordered domain, or passage through partially unfolded states that enables them to morph into an eclectic menagerie of misfolded structures with variable toxicities [15]. These structures include self-templating amyloid fibrils with cross-β architecture, disordered aggregates, and small soluble oligomers [1,3]. For example, in PD, a small intrinsically-disordered protein, α-syn, forms amyloid fibrils that self-template or ‘seed’ their own assembly via recruiting soluble forms of α-syn to their elongating ends where α-syn is conformationally converted to the cross-β structure [1,68,9•,10•]. α-Syn amyloid can spread from cell to cell, thereby propagating pathology [1,6,8,11,12••,13,14]. Indeed, amyloid fibrils formed by recombinant α-syn in the test tube can induce a PD-like disease when injected into the brain of a mouse [6,11,13]. This transforming principle establishes that the self-replicating structure of α-syn amyloid can encode the PD phenotype, which develops via the ongoing conversion of endogenous α-syn to the amyloid state as α-syn fibrils spread through the brain [1,8,13,15]. Moreover, α-syn can form fibrils with different cross-β structures, termed ‘strains’, which encode distinct neurodegenerative phenotypes [9•,10•,16,17•,18,19]. The lateral face of α-syn amyloid provides a surface where α-syn oligomers can nucleate [20]. α-Syn populates diverse soluble, oligomeric species before, during, and after α-syn amyloidogenesis, which can be on or off pathway for amyloid formation [7,2125,26•]. α-Syn oligomers are typically more toxic than mature fibrils [7,23]. Small soluble oligomers or short, fragmented amyloid fibrils are more toxic than very large aggregated species, which due to their low surface-area-to-volume-ratio shield damaging surfaces inside the aggregate [7,23,27]. A major challenge for any therapeutic aimed at mitigating protein misfolding is the ability to remodel diverse, toxic misfolded conformers, including soluble oligomers and amyloid fibrils into benign species [28,29].

Protein disaggregases as potential therapeutics

I have postulated that protein disaggregases could be uniquely suited to meet this challenge as they can safely deconstruct self-templating amyloid and toxic soluble oligomers, and recover soluble protein with restored functionality from these structures [28,29]. Thus, protein disaggregases could mitigate any toxic gain-of-function or toxic loss-of-function connected with protein misfolding, and simultaneously could eradicate self-templating species that propagate disease [28,29]. Protein disaggregation might also be coupled to protein degradation, which could also be beneficial to eliminate toxic and self-templating conformers, and subsequent translation of new protein could antagonize any toxic loss-of-function [29]. However, protein disaggregases remain among the least understood components of the proteostasis network, and we are only at the inception of realizing their existence and potential [28,29]. Here, I highlight recent advances to define, engineer, and apply protein disaggregases to reverse deleterious protein misfolding in neurodegenerative disease.

Hsp104, a protein disaggregase from yeast

Hsp104 is an asymmetric ring-shaped translocase and hexameric AAA+ protein found in yeast [30,31••]. Hsp104 couples ATP hydrolysis to the rapid dissolution and reactivation of diverse proteins trapped in disordered aggregates, ordered stress-induced assemblies, preamyloid oligomers, amyloids, and prions [3237,38••]. Optimal Hsp104 disaggregase activity can require collaboration with the Hsp70 chaperone system [30]. In yeast, Hsp104 performs critical functions in stress tolerance, prion inheritance, asymmetric partitioning of aggregates during cell division, and promoting longevity [30]. Hsp104 rapidly disaggregates Sup35 prions within a few minutes [3537,39]. Moreover, Hsp104 effectively dissolves amyloids formed by diverse human degenerative disease proteins, including: Aβ42, tau, polyglutamine, α-syn, prion protein (PrP), and amylin [34,4042]. Hsp104 also rapidly remodels amyloid fibrils formed by fragments of prostatic acid phosphatase (PAP248-286 and PAP85-120) [43••], which are abundant in human seminal fluid and promote HIV infection [44]. This rapid and broad-spectrum amyloid-disaggregase activity of Hsp104 is unusual and might represent a therapeutic opportunity [28].

Intriguingly, Hsp104 is absent from metazoa, but is found in all non-metazoan eukaryotes, all eubacteria, and some archaebacteria [45]. Thus, Hsp104 could be developed into a vital disruptive technology that retools proteostasis to combat neurodegenerative disease and HIV infection [28,43••,46]. Indeed, we have established Hsp104 as the only factor known to dissociate α-syn oligomers and amyloids connected with PD and rescue α-syn-induced neurodegeneration in the substantia nigra of a rat PD model [34,40,41]. Moreover, Hsp104 rescues polyglutamine toxicity and neurodegeneration in C. elegans, fly, mouse, and rat [28,47]. Hsp104 even rescues polyglutamine toxicity after degeneration has begun [47]. Hsp104 expression is not detrimental in metazoa and can be broadly and safely expressed in worm, fly, mouse, and rat, as well as in mammalian cell and neuronal cultures [28,34,47]. These findings make it difficult to understand why Hsp104 was lost from metazoa, but also emphasize that Hsp104 might be safely introduced and developed as a therapeutic agent [28,43••,46].

Despite these encouraging activities, very high Hsp104 concentrations are needed for optimal disaggregation of human disease proteins, such as α-syn, which may restrict efficacy [34,40,41]. Thus, we have engineered potentiated Hsp104 variants, which rescue aggregation and toxicity of proteins associated with neurodegenerative disease such as TDP-43, FUS, TAF15, and α-syn, and mitigate neurodegeneration in the metazoan nervous system at concentrations where Hsp104 is ineffective [46,48••,49,50,51••,52]. Hsp104 activity can be potentiated by single missense mutations at specific positions in the middle domain or nucleotide-binding domain 1 of Hsp104 [46]. Potentiating mutations reconfigure how Hsp104 subunits collaborate, alter substrate discrimination, alleviate any stringent requirements for Hsp70, and enhance Hsp104’s ATPase, translocase (rate at which substrates are translocated across the central channel of Hsp104), unfoldase, and disaggregase activity [48••,49,50]. These combined properties enable potentiated Hsp104 variants to outperform Hsp104 under conditions where an aggregation-prone protein such as TDP-43, FUS, or α-syn has exceeded proteostatic buffers and is undergoing widespread misfolding and aggregation [46]. Potentiated Hsp104 variants can have off-target effects [48••,49], and may require further engineering to minimize these via increasing substrate selectivity [46]. Importantly, substoichiometric concentrations of potentiated Hsp104 variants can remodel amyloid [43••]. For example, nanomolar concentrations of an enhanced Hsp104 variant, Hsp104A503V, can remodel micromolar concentrations of PAP248-296 sequestered in SEVI fibrils [43••]. The challenge ahead is to determine whether Hsp104 or enhanced variants can confer increased therapeutic benefits in mammalian cells, patient-derived neurons, and additional animal models of neurodegeneration.

An issue that is often raised about introducing any exogenous protein as a therapeutic is whether the patient might mount a deleterious immune response against the therapeutic protein. However, it is important to note that the central nervous system (CNS) exhibits immune privilege [53]. Thus, immune responses to CNS antigens are very slow to develop [53], which could provide a therapeutic window for exogenous agents, such as Hsp104 [29], tetanus and botulinum toxin variants [54], or even CRISPR-Cas9 [55] delivered to the CNS. Indeed, we observed no deleterious side effects of expressing Hsp104 for 6 weeks in the rat substantia nigra when delivered via lentivirus [34]. Transient expression or delivery of an exogenous therapeutic protein to the CNS may thus be well tolerated (particularly if combined with an immunosuppressant), although significant caution is highly warranted.

Gene 3 protein, a protein disaggregase from bacteriophage M13

Hsp104 must hydrolyze ATP to disaggregate proteins [30], which would limit disaggregase activity in ATP-depleted environments such as the extracellular space, where the ATP concentration is ~10nM compared to ~3–10mM inside cells [56]. The extracellular space is where Aβ deposits accumulate in AD and prions accumulate in Creutzfeldt-Jakob Disease [1]. Thus, to antagonize these extracellular protein-misfolding events Hsp104 would need to be engineered to operate effectively at limiting ATP concentrations. An alternative strategy is to define ATP-independent protein disaggregases, which may couple binding energy to disaggregation and are operational outside cells. Select small molecules, including CLR01, a lysine- and arginine-specific molecular tweezer, and the green tea polyphenol, EGCG, can safely disaggregate or remodel diverse amyloids [57,58,59••,60]. A human Aβ oligomer- and fibril-specific monoclonal antibody, aducanumab, promotes clearance of Aβ plaques and may retard clinical decline in AD patients and is now in phase 3 trials [61••]. However, whether this Aβ-plaque clearance is due to fibril disaggregation (as with select anti-Aβ antibodies that bind the N-terminal region of Aβ [62,63]), phagocytic clearance by microglia, or both is uncertain [61••]. Several other ATP-independent protein disaggregases have emerged including a subunit of the chloroplast signal recognition particle [6466], cyclophilin [6769], HtrA1 [70••], and gene 3 protein (g3p) [71••].

G3p is a minor capsid protein from filamentous bacteriophage M13, which enables viral entry into the bacterial host [72]. Remarkably, g3p enables M13 bacteriophages to slowly remodel diverse amyloids, including those formed by Aβ42, α-syn, tau, and NM (the prion domain of Sup35) [71••,73]. This amyloid-remodeling activity may enable phages to infect bacteria by penetrating protective amyloid-based biofilms [74]. The amyloid-binding activity of g3p was mapped to its two N-terminal domains, which are separated by a glycine-rich hinge [71••]. A recombinant soluble g3p fragment, termed G3P, which comprises the two N-terminal domains and the hinge could bind diverse amyloids but could not disaggregate them [71••]. However, 3–5 copies of g3p form an oligomer at the tip of the filamentous phage capsid [72]. Thus, g3p multivalency may enable amyloid remodeling [71••]. Indeed, amyloid-disaggregation activity was restored by engineering G3P to be dimeric via an immunoglobulin (Ig) Fc-G3P fusion protein (Ig-G3P) [71••]. Importantly, Ig-G3P selectively bound and disaggregated diverse amyloids, including Aβ42 and tau fibrils, and did not interact with various disordered aggregates or monomeric Aβ42 [71••]. It is not yet clear whether Ig-G3P remodels soluble toxic oligomers. Although the mechanism by which Ig-G3P dissociates diverse amyloids is uncertain, multiple binding events to different regions of assembled fibrils seems likely to be important [71••]. Importantly, weekly intraperitoneal injection of Ig-G3P reduced both Aβ and tau pathologies and improved cognition in mouse models [75•]. Thus, Ig-G3P is an interesting therapeutic candidate for AD that targets Aβ and tau misfolding, and is now in Phase 1B clinical trials [75•].

Hsp110, Hsp70, and Hsp40 disaggregases in humans

Bafflingly, Hsp104 is absent from metazoa [28,45], and whether metazoa even possess a protein disaggregation and reactivation machinery had endured as a long-standing enigma [41,52]. It is now clear that human Hsp110, Hsp70, and Hsp40 synergize to dissolve and reactivate model proteins trapped in disordered aggregates and depolymerize amyloid fibrils formed by α-syn [41,52,7680]. Small heat-shock proteins can further enhance the disaggregase activity of this system [76]. Proteins disaggregated by Hsp110, Hsp70, and Hsp40 can be refolded [41,80] or passed to the proteasome to be degraded [81••]. This latter pathway is mediated by Ubiquilin 2, which recognizes client-bound Hsp70 and enables transfer of client to the proteasome [81••]. Interestingly, ALS-linked mutations in Ubiquilin 2 impair this activity, which may contribute to disease [81••].

Hsp110, Hsp70, and Hsp40 drive disaggregation by exerting pulling forces on aggregated polypeptides, which are forcibly extracted from the aggregate [52,82,83]. This system might also remodel toxic soluble oligomers formed by various proteins [37,84]. Hsp70 must engage substrate and Hsp110, and hydrolyze ATP to drive protein disaggregation [41,52]. Hsp40 must harbor a functional J domain (which stimulates Hsp110 and Hsp70 ATPase activity) to promote protein disaggregation, but the J domain alone is insufficient [41,52]. Whether Hsp110 acts simply as a nucleotide exchange factor (NEF) for Hsp70 or whether it must also bind substrate, hydrolyze ATP, or both as part of the disaggregation reaction is debated [52,82,83]. Likewise, whether Hsp70 must act as a NEF for Hsp110 to drive protein disaggregation is debated [52,78,83]. I suspect there is plasticity in disaggregase mechanism with respect to the exact role of Hsp110, which may depend on aggregate structure as with Hsp104 [40,52]. Indeed, Hsp70 (and likely Hsp110) exhibits functional plasticity via alternative modes of client engagement, which can promote protein folding or unfolding [85•]. Regardless, humans express a variety of Hsp110, Hsp70, and Hsp40 chaperones, and the precise combination and ratio of components can enhance or inhibit disaggregase activity against different substrates [41,7780].

The Hsp110, Hsp70, and Hsp40 disaggregase machinery must become overwhelmed in neurodegenerative disease. Indeed, Hsp110 knockout mice develop age-dependent tau hyperphosphorylation, early accumulation of insoluble Aβ, and neurodegeneration [86]. Moreover, Hsp110 (or another Hsp70 NEF, Bag3), Hsp70, Hsp40, and small heat-shock proteins collaborate to dissolve stress granules [87,88,89••,90], dynamic RNP assemblies that accumulate upon stress (and incorporate TDP-43 and FUS), and are connected to ALS and formation of pathological aggregates [2,4]. Upregulation or stimulation of Hsp110, Hsp70, and Hsp40 disaggregase activity, perhaps with small-molecule drugs [91], could have key therapeutic applications [29]. Importantly, Hsp104 can greatly enhance the disaggregase activity of Hsp110, Hsp70, and Hsp40 [41,76]. Co-expression of Hsp110 and Hsp40 in Drosophila suppresses polyglutamine toxicity [92]. Moreover, increased expression of Hsp110 extends lifespan of ALS-linked SOD1 transgenic mice [93••].

Engineering Hsp110, Hsp70, and Hsp40 to have enhanced disaggregase activity may enable robust neuroprotection [94]. Hsp70 variants with enhanced chaperone or disaggregase activity against specific model substrates have been uncovered [9497], but whether these can be translated into neuroprotective agents in vivo is unknown. Engineering a secreted form of Hsp70 rescued a Drosophila model of AD, although in this case Hsp70 promoted clustering of Aβ42 into larger aggregates with reduced neurotoxicity [98•]. This clustering activity appears to be a general feature of molecular chaperones that must operate under ATP-limited conditions such as the extracellular space, and can be protective by minimizing exposure of reactive surfaces by confining them within the aggregate interior [43••,99].

HtrA1, an ATP-independent human protein disaggregase

HtrA1 is a ubiquitously expressed chaperone and homo-oligomeric PDZ serine protease abundant in human brain [100,101], which is also an ATP-independent protein disaggregase [70••]. HtrA1 is localized to the cytoplasm and extracellular space, and selectively degrades misfolded substrates while leaving their folded counterparts alone [29,70••,100,101]. HtrA1 disassembles and degrades tau and Aβ42 fibrils connected to AD [70••]. An engineered protease-defective HtrA1 variant dissolves tau and Aβ42 fibrils without degrading them [70••]. Thus, HtrA1 could be tailored to dissolve inclusions in AD or tauopathies, which could be important to rapidly restore tau loss-of-function [29]. A significant inverse correlation exists between HtrA1 and tau levels in AD patient brains [101]. These data suggest that HtrA1 functions as a tau disaggregase and protease in vivo [70••,101]. Whether HtrA1 can dissociate and degrade toxic soluble oligomers is uncertain. Nonetheless, HtrA1 could be a valuable ATP-independent protein disaggregase against AD, and like Ig-G3P targets both tau and Aβ42 [70••,71••]. It will be important to test whether elevating HtrA1 activity is protective in animal models of AD and tauopathy. Intriguingly, defects in the mitochondrial isoform, HtrA2, have been connected to PD [102].

NMNAT2 and Hsp90 combine to refold aggregated proteins

Nicotinamide mononucleotide adenylyl transferases (NMNATs) synthesize nicotinamide adenine dinucleotide (NAD+) [103], a critical co-enzyme that performs key electron-transfer events in metabolism and is an essential substrate for sirtuins and poly(adenosine diphosphate-ribose) polymerases [104]. Humans express three NMNATs, with NMNAT2 being abundant in the brain, whereas Drosophila express a single NMNAT [103]. Remarkably, NMNATs can be neuroprotective in several models of neurodegenerative disease, and can function as chaperones that prevent aggregation of disease proteins, including polyglutamine-expanded ataxin 1 and tau [103,105,106]. NMNAT2 collaborates with Hsp90 to disaggregate and refold previously aggregated proteins [107••]. NMNAT2 disaggregase activity is independent from NAD+ biosynthesis, but requires a unique C-terminal ATP-binding site that may be activated by Hsp90 [107••]. Thus, NMNAT2 chaperone and disaggregase activity may reduce proteotoxicity, whereas its NAD+ biosynthetic activity may protect neurons from excitotoxicity [107••]. Indeed, NAD+ replenishment strategies may confer neuroprotection in prion diseases [108]. In the absence of NMNAT2, Hsp90 can depolymerize TDP-43 fibrils in vitro [109]. The precise mechanism by which NMNAT2 and Hsp90 combine to promote protein disaggregation and reactivation remains unclear. Nonetheless, methods to stimulate the disaggregase activity or upregulate expression of NMNAT2, Hsp90, or both, perhaps with small-molecule drugs, could be important to combat neurodegenerative disease.

Conclusions and Future Directions

Endogenous human protein disaggregases fail to counter neurodegenerative disease, perhaps due to reduced expression or activity in selectively vulnerable neurons [93••,101,107••,110]. Thus, augmenting or stimulating their activity could have therapeutic utility [29]. Here, I have highlighted several natural and engineered protein-disaggregase modalities that could be appropriated for therapeutic purposes. Excitingly, Ig-G3P is now in clinical trials for AD, and Hsp104 [34,37,47,48••], Hsp110, Hsp70, and Hsp40 [92,93••], and NMNAT2 plus Hsp90 [106,107••] have shown efficacy in animal models of neurodegenerative disease. Additional fine tuning of disaggregase activity for specific substrates may help optimize each system for specific disorders [29]. There is also great interest in defining small-molecule drugs that increase expression or directly enhance the activity of endogenous human protein disaggregases [29]. It will also be important to determine whether natural polymorphisms in endogenous human molecular chaperones or protein disaggregases [111] enhance their activity and render individuals more resistant to developing a neurodegenerative disease. We are still only beginning to realize the existence and therapeutic potential of protein disaggregases and many challenges lie ahead in translation to therapeutics [29]. Nonetheless, protein disaggregases represent a valuable opportunity to develop treatments for several devastating neurodegenerative diseases.

Highlights.

  • Showcases advances to define, engineer, and apply protein disaggregases to mitigate neurodegenerative disease associated with protein misfolding.

  • Highlights therapeutic disaggregase activity of Hsp104 and directed engineered variants with robust activity against clients linked to neurodegeneration.

  • Highlights therapeutic disaggregase activity of gene 3 protein and engineered variants with robust activity against clients linked to neurodegeneration.

  • Detailed discussion of the disaggregase activities of human Hsp110, Hsp70 and Hsp40; human HtrA1; and human NMNAT2 with Hsp90.

  • Emphasis on how engineered variants of powerful chaperones with newly identified disaggregase activity can be harnessed against clients linked to neurodegeneration.

Acknowledgments

Thanks to Korrie Mack, Zachary March, and Edward Chuang for feedback on the manuscript. J.S. is supported by the NIH (R01GM099836 and R21NS090205), Life Extension Foundation, a Sanofi Innovation award, ALS Association, Muscular Dystrophy Association, Target ALS, and the Robert Packard Center for ALS Research at Johns Hopkins.

Footnotes

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

References

  • 1.Cushman M, Johnson BS, King OD, Gitler AD, Shorter J. Prion-like disorders: blurring the divide between transmissibility and infectivity. J Cell Sci. 2010;123:1191–1201. doi: 10.1242/jcs.051672. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 2.March ZM, King OD, Shorter J. Prion-like domains as epigenetic regulators, scaffolds for subcellular organization, and drivers of neurodegenerative disease. Brain Res. 2016;1647:9–18. doi: 10.1016/j.brainres.2016.02.037. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 3.Eisele YS, Monteiro C, Fearns C, Encalada SE, Wiseman RL, Powers ET, Kelly JW. Targeting protein aggregation for the treatment of degenerative diseases. Nat Rev Drug Discov. 2015;14:759–780. doi: 10.1038/nrd4593. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 4.Li YR, King OD, Shorter J, Gitler AD. Stress granules as crucibles of ALS pathogenesis. J Cell Biol. 2013;201:361–372. doi: 10.1083/jcb.201302044. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 5.Uversky VN, Fink AL. Conformational constraints for amyloid fibrillation: the importance of being unfolded. Biochim Biophys Acta. 2004;1698:131–153. doi: 10.1016/j.bbapap.2003.12.008. [DOI] [PubMed] [Google Scholar]
  • 6.Hasegawa M, Nonaka T, Masuda-Suzukake M. alpha-Synuclein: Experimental Pathology. Cold Spring Harb Perspect Med. 2016:6. doi: 10.1101/cshperspect.a024273. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 7.Lashuel HA, Overk CR, Oueslati A, Masliah E. The many faces of alpha-synuclein: from structure and toxicity to therapeutic target. Nat Rev Neurosci. 2013;14:38–48. doi: 10.1038/nrn3406. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 8.Luna E, Luk KC. Bent out of shape: alpha-Synuclein misfolding and the convergence of pathogenic pathways in Parkinson’s disease. FEBS Lett. 2015;589:3749–3759. doi: 10.1016/j.febslet.2015.10.023. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • •9.Rodriguez JA, Ivanova MI, Sawaya MR, Cascio D, Reyes FE, Shi D, Sangwan S, Guenther EL, Johnson LM, Zhang M, et al. Structure of the toxic core of alpha-synuclein from invisible crystals. Nature. 2015;525:486–490. doi: 10.1038/nature15368. The authors use micro-electron diffraction to determine the structure of amyloid fibrils formed by an 11-residue segment of human α-synuclein at atomic resolution, which reveals new opportunities to rationally design inhibitors of α-synuclein amyloidogenesis. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • •10.Tuttle MD, Comellas G, Nieuwkoop AJ, Covell DJ, Berthold DA, Kloepper KD, Courtney JM, Kim JK, Barclay AM, Kendall A, et al. Solid-state NMR structure of a pathogenic fibril of full-length human alpha-synuclein. Nat Struct Mol Biol. 2016;23:409–415. doi: 10.1038/nsmb.3194. The authors use solid-state NMR spectroscopy, EM, and X-ray fiber diffraction to solve a high-resolution structure of an α-synuclein fibril that induces Lewy body pathology in primary neurons. The fibril adopts a novel orthogonal Greek-key topology that enables robust self-templating activity. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 11.Luk KC, Kehm V, Carroll J, Zhang B, O’Brien P, Trojanowski JQ, Lee VM. Pathological alpha-synuclein transmission initiates Parkinson-like neurodegeneration in nontransgenic mice. Science. 2012;338:949–953. doi: 10.1126/science.1227157. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • ••12.Mao X, Ou MT, Karuppagounder SS, Kam TI, Yin X, Xiong Y, Ge P, Umanah GE, Brahmachari S, Shin JH, et al. Pathological alpha-synuclein transmission initiated by binding lymphocyte-activation gene 3. Science. 2016:353. doi: 10.1126/science.aah3374. This study establishes that α-synuclein fibrils bind to LAG3 (lymphocyte-activation gene 3), which enables α-synuclein fibrils to spread through the brain and cause toxicity. Deletion of LAG3 retarded the ability of α-synuclein fibrils to induce dopaminergic neurodegeneration in vivo. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 13.Uchihara T, Giasson BI. Propagation of alpha-synuclein pathology: hypotheses, discoveries, and yet unresolved questions from experimental and human brain studies. Acta Neuropathol. 2016;131:49–73. doi: 10.1007/s00401-015-1485-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14.Volpicelli-Daley LA, Luk KC, Patel TP, Tanik SA, Riddle DM, Stieber A, Meaney DF, Trojanowski JQ, Lee VM. Exogenous alpha-synuclein fibrils induce Lewy body pathology leading to synaptic dysfunction and neuron death. Neuron. 2011;72:57–71. doi: 10.1016/j.neuron.2011.08.033. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 15.Mahul-Mellier AL, Vercruysse F, Maco B, Ait-Bouziad N, De Roo M, Muller D, Lashuel HA. Fibril growth and seeding capacity play key roles in alpha-synuclein-mediated apoptotic cell death. Cell Death Differ. 2015;22:2107–2122. doi: 10.1038/cdd.2015.79. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16.Guo JL, Covell DJ, Daniels JP, Iba M, Stieber A, Zhang B, Riddle DM, Kwong LK, Xu Y, Trojanowski JQ, et al. Distinct alpha-synuclein strains differentially promote tau inclusions in neurons. Cell. 2013;154:103–117. doi: 10.1016/j.cell.2013.05.057. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • •17.Prusiner SB, Woerman AL, Mordes DA, Watts JC, Rampersaud R, Berry DB, Patel S, Oehler A, Lowe JK, Kravitz SN, et al. Evidence for alpha-synuclein prions causing multiple system atrophy in humans with parkinsonism. Proc Natl Acad Sci U S A. 2015;112:E5308–5317. doi: 10.1073/pnas.1514475112. In this study, evidence is presented that multiple system atrophy (MSA) is caused by a unique strain of α-synuclein prions, which is distinct from the putative α-synuclein prions that cause PD. MSA may then represent a new prion disease. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 18.Woerman AL, Stohr J, Aoyagi A, Rampersaud R, Krejciova Z, Watts JC, Ohyama T, Patel S, Widjaja K, Oehler A, et al. Propagation of prions causing synucleinopathies in cultured cells. Proc Natl Acad Sci U S A. 2015;112:E4949–4958. doi: 10.1073/pnas.1513426112. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 19.Peelaerts W, Bousset L, Van der Perren A, Moskalyuk A, Pulizzi R, Giugliano M, Van den Haute C, Melki R, Baekelandt V. alpha-Synuclein strains cause distinct synucleinopathies after local and systemic administration. Nature. 2015;522:340–344. doi: 10.1038/nature14547. [DOI] [PubMed] [Google Scholar]
  • 20.Buell AK, Galvagnion C, Gaspar R, Sparr E, Vendruscolo M, Knowles TP, Linse S, Dobson CM. Solution conditions determine the relative importance of nucleation and growth processes in alpha-synuclein aggregation. Proc Natl Acad Sci U S A. 2014;111:7671–7676. doi: 10.1073/pnas.1315346111. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21.Conway KA, Rochet JC, Bieganski RM, Lansbury PT., Jr Kinetic stabilization of the alpha-synuclein protofibril by a dopamine-alpha-synuclein adduct. Science. 2001;294:1346–1349. doi: 10.1126/science.1063522. [DOI] [PubMed] [Google Scholar]
  • 22.Diogenes MJ, Dias RB, Rombo DM, Vicente Miranda H, Maiolino F, Guerreiro P, Nasstrom T, Franquelim HG, Oliveira LM, Castanho MA, et al. Extracellular alpha-synuclein oligomers modulate synaptic transmission and impair LTP via NMDA-receptor activation. J Neurosci. 2012;32:11750–11762. doi: 10.1523/JNEUROSCI.0234-12.2012. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 23.Kayed R, Head E, Thompson JL, McIntire TM, Milton SC, Cotman CW, Glabe CG. Common structure of soluble amyloid oligomers implies common mechanism of pathogenesis. Science. 2003;300:486–489. doi: 10.1126/science.1079469. [DOI] [PubMed] [Google Scholar]
  • 24.Lashuel HA, Hartley D, Petre BM, Walz T, Lansbury PT., Jr Neurodegenerative disease: amyloid pores from pathogenic mutations. Nature. 2002;418:291. doi: 10.1038/418291a. [DOI] [PubMed] [Google Scholar]
  • 25.Lorenzen N, Nielsen SB, Buell AK, Kaspersen JD, Arosio P, Vad BS, Paslawski W, Christiansen G, Valnickova-Hansen Z, Andreasen M, et al. The role of stable alpha-synuclein oligomers in the molecular events underlying amyloid formation. J Am Chem Soc. 2014;136:3859–3868. doi: 10.1021/ja411577t. [DOI] [PubMed] [Google Scholar]
  • •26.Paslawski W, Mysling S, Thomsen K, Jorgensen TJ, Otzen DE. Co-existence of two different alpha-synuclein oligomers with different core structures determined by hydrogen/deuterium exchange mass spectrometry. Angew Chem Int Ed Engl. 2014;53:7560–7563. doi: 10.1002/anie.201400491. In this study, hydrogen/deuterium exchange monitored by mass spectrometry is used to define structural features of two different types of soluble α-synuclein oligomers. The first type evolves into α-synuclein fibrils, while the second transforms into amorphous aggregates. [DOI] [PubMed] [Google Scholar]
  • 27.Xue WF, Hellewell AL, Gosal WS, Homans SW, Hewitt EW, Radford SE. Fibril fragmentation enhances amyloid cytotoxicity. J Biol Chem. 2009;284:34272–34282. doi: 10.1074/jbc.M109.049809. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28.Shorter J. Hsp104: a weapon to combat diverse neurodegenerative disorders. Neurosignals. 2008;16:63–74. doi: 10.1159/000109760. [DOI] [PubMed] [Google Scholar]
  • 29.Shorter J. Engineering therapeutic protein disaggregases. Mol Biol Cell. 2016;27:1556–1560. doi: 10.1091/mbc.E15-10-0693. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30.Sweeny EA, Shorter J. Mechanistic and Structural Insights into the Prion-Disaggregase Activity of Hsp104. J Mol Biol. 2016;428:1870–1885. doi: 10.1016/j.jmb.2015.11.016. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • ••31.Yokom AL, Gates SN, Jackrel ME, Mack KL, Su M, Shorter J, Southworth DR. Spiral architecture of the Hsp104 disaggregase reveals the basis for polypeptide translocation. Nat Struct Mol Biol. 2016;23:830–837. doi: 10.1038/nsmb.3277. In this study, the highest resolution structure of the Hsp104 hexamer to date is solved by cryo-EM, which reveals a left-handed spiral maintained by an unprecedented heteromeric AAA+ interaction. Substrate-binding regions line the central Hsp104 channel in a spiral arrangement, which provide a continuous path for polypeptide translocation during disaggregation. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 32.Glover JR, Lindquist S. Hsp104, Hsp70, and Hsp40: a novel chaperone system that rescues previously aggregated proteins. Cell. 1998;94:73–82. doi: 10.1016/s0092-8674(00)81223-4. [DOI] [PubMed] [Google Scholar]
  • 33.Klaips CL, Hochstrasser ML, Langlois CR, Serio TR. Spatial quality control bypasses cell-based limitations on proteostasis to promote prion curing. Elife. 2014:3. doi: 10.7554/eLife.04288. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 34.Lo Bianco C, Shorter J, Regulier E, Lashuel H, Iwatsubo T, Lindquist S, Aebischer P. Hsp104 antagonizes alpha-synuclein aggregation and reduces dopaminergic degeneration in a rat model of Parkinson disease. J Clin Invest. 2008;118:3087–3097. doi: 10.1172/JCI35781. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 35.Shorter J, Lindquist S. Hsp104 catalyzes formation and elimination of self-replicating Sup35 prion conformers. Science. 2004;304:1793–1797. doi: 10.1126/science.1098007. [DOI] [PubMed] [Google Scholar]
  • 36.Shorter J, Lindquist S. Destruction or potentiation of different prions catalyzed by similar Hsp104 remodeling activities. Mol Cell. 2006;23:425–438. doi: 10.1016/j.molcel.2006.05.042. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37.Shorter J, Lindquist S. Hsp104, Hsp70 and Hsp40 interplay regulates formation, growth and elimination of Sup35 prions. EMBO J. 2008;27:2712–2724. doi: 10.1038/emboj.2008.194. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • ••38.Wallace EW, Kear-Scott JL, Pilipenko EV, Schwartz MH, Laskowski PR, Rojek AE, Katanski CD, Riback JA, Dion MF, Franks AM, et al. Reversible, Specific, Active Aggregates of Endogenous Proteins Assemble upon Heat Stress. Cell. 2015;162:1286–1298. doi: 10.1016/j.cell.2015.08.041. In this study, a mild heat shock in yeast is shown to induce ordered protein assemblies that do not contain grossly misfolded proteins. After stress, these structures are readily resolved confirming that proteins are reactivated and not degraded after a mild heat shock in yeast. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 39.DeSantis ME, Shorter J. Hsp104 drives “protein-only” positive selection of Sup35 prion strains encoding strong [PSI(+)] Chem Biol. 2012;19:1400–1410. doi: 10.1016/j.chembiol.2012.09.013. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 40.DeSantis ME, Leung EH, Sweeny EA, Jackrel ME, Cushman-Nick M, Neuhaus-Follini A, Vashist S, Sochor MA, Knight MN, Shorter J. Operational plasticity enables Hsp104 to disaggregate diverse amyloid and nonamyloid clients. Cell. 2012;151:778–793. doi: 10.1016/j.cell.2012.09.038. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 41.Shorter J. The mammalian disaggregase machinery: Hsp110 synergizes with Hsp70 and Hsp40 to catalyze protein disaggregation and reactivation in a cell-free system. PLoS One. 2011;6:e26319. doi: 10.1371/journal.pone.0026319. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 42.Liu YH, Han YL, Song J, Wang Y, Jing YY, Shi Q, Tian C, Wang ZY, Li CP, Han J, et al. Heat shock protein 104 inhibited the fibrillization of prion peptide 106–126 and disassembled prion peptide 106–126 fibrils in vitro. Int J Biochem Cell Biol. 2011;43:768–774. doi: 10.1016/j.biocel.2011.01.022. [DOI] [PubMed] [Google Scholar]
  • ••43.Castellano LM, Bart SM, Holmes VM, Weissman D, Shorter J. Repurposing Hsp104 to Antagonize Seminal Amyloid and Counter HIV Infection. Chem Biol. 2015;22:1074–1086. doi: 10.1016/j.chembiol.2015.07.007. In this study, Hsp104 and a potentiated Hsp104 variant are shown to remodel various natural seminal fluid amyloids that promote HIV infection. By engineering Hsp104 to interact with the chambered protease, ClpP, it is shown that Hsp104 can couple amyloid remodeling to degradation. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 44.Castellano LM, Shorter J. The surprising role of amyloid fibrils in HIV infection. Biology. 2012;1:58–80. doi: 10.3390/biology1010058. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45.Erives AJ, Fassler JS. Metabolic and chaperone gene loss marks the origin of animals: evidence for Hsp104 and Hsp78 chaperones sharing mitochondrial enzymes as clients. PLoS One. 2015;10:e0117192. doi: 10.1371/journal.pone.0117192. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46.Jackrel ME, Shorter J. Engineering enhanced protein disaggregases for neurodegenerative disease. Prion. 2015;9:90–109. doi: 10.1080/19336896.2015.1020277. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 47.Cushman-Nick M, Bonini NM, Shorter J. Hsp104 suppresses polyglutamine-induced degeneration post onset in a Drosophila MJD/SCA3 model. PLoS Genet. 2013;9:e1003781. doi: 10.1371/journal.pgen.1003781. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • ••48.Jackrel ME, DeSantis ME, Martinez BA, Castellano LM, Stewart RM, Caldwell KA, Caldwell GA, Shorter J. Potentiated Hsp104 variants antagonize diverse proteotoxic misfolding events. Cell. 2014;156:170–182. doi: 10.1016/j.cell.2013.11.047. In this study, Hsp104 variants are engineered with enhanced disaggregase activity against TDP-43, FUS, and α-synuclein. This work establishes that disease-associated aggregates and amyloid are tractable targets and that potentiated disaggregases can restore proteostasis and mitigate neurodegeneration. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 49.Jackrel ME, Shorter J. Potentiated Hsp104 variants suppress toxicity of diverse neurodegenerative disease-linked proteins. Dis Model Mech. 2014;7:1175–1184. doi: 10.1242/dmm.016113. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 50.Jackrel ME, Yee K, Tariq A, Chen AI, Shorter J. Disparate Mutations Confer Therapeutic Gain of Hsp104 Function. ACS Chem Biol. 2015;10:2672–2679. doi: 10.1021/acschembio.5b00765. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • ••51.Sweeny EA, Jackrel ME, Go MS, Sochor MA, Razzo BM, DeSantis ME, Gupta K, Shorter J. The Hsp104 N-terminal domain enables disaggregase plasticity and potentiation. Mol Cell. 2015;57:836–849. doi: 10.1016/j.molcel.2014.12.021. In this study, the precise mechanism by which Hsp104 dissolves Sup35 prions is delineated. The N-terminal domain of Hsp104 plays a critical role, which enables Hsp104 to dissolve Sup35 prions and not merely fragment them. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 52.Torrente MP, Shorter J. The metazoan protein disaggregase and amyloid depolymerase system: Hsp110, Hsp70, Hsp40, and small heat shock proteins. Prion. 2013;7:457–463. doi: 10.4161/pri.27531. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 53.Louveau A, Harris TH, Kipnis J. Revisiting the Mechanisms of CNS Immune Privilege. Trends Immunol. 2015;36:569–577. doi: 10.1016/j.it.2015.08.006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 54.Ovsepian SV, O’Leary VB, Ntziachristos V, Dolly JO. Circumventing Brain Barriers: Nanovehicles for Retroaxonal Therapeutic Delivery. Trends Mol Med. 2016;22:983–993. doi: 10.1016/j.molmed.2016.09.004. [DOI] [PubMed] [Google Scholar]
  • 55.McMahon MA, Cleveland DW. Gene therapy: Gene-editing therapy for neurological disease. Nat Rev Neurol. 2017;13:7–9. doi: 10.1038/nrneurol.2016.190. [DOI] [PubMed] [Google Scholar]
  • 56.Trautmann A. Extracellular ATP in the immune system: more than just a “danger signal”. Sci Signal. 2009;2:pe6. doi: 10.1126/scisignal.256pe6. [DOI] [PubMed] [Google Scholar]
  • 57.Shorter J. Emergence and natural selection of drug-resistant prions. Mol Biosyst. 2010;6:1115–1130. doi: 10.1039/c004550k. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 58.Schrader T, Bitan G, Klarner FG. Molecular tweezers for lysine and arginine - powerful inhibitors of pathologic protein aggregation. Chem Commun (Camb) 2016;52:11318–11334. doi: 10.1039/c6cc04640a. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • ••59.Lump E, Castellano LM, Meier C, Seeliger J, Erwin N, Sperlich B, Sturzel CM, Usmani S, Hammond RM, von Einem J, et al. A molecular tweezer antagonizes seminal amyloids and HIV infection. Elife. 2015:4. doi: 10.7554/eLife.05397. In this study, a lysine- and arginine-specific molecular tweezer, CLR01, is shown to disrupt seminal amyloids that promote HIV infection. Remarkably, CLR01 also disrupts diverse enveloped viruses and could be a promising topical microbicide. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 60.Castellano LM, Hammond RM, Holmes VM, Weissman D, Shorter J. Epigallocatechin-3-gallate rapidly remodels PAP85-120, SEM1(45-107), and SEM2(49-107) seminal amyloid fibrils. Biol Open. 2015;4:1206–1212. doi: 10.1242/bio.010215. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • ••61.Sevigny J, Chiao P, Bussiere T, Weinreb PH, Williams L, Maier M, Dunstan R, Salloway S, Chen T, Ling Y, et al. The antibody aducanumab reduces Abeta plaques in Alzheimer’s disease. Nature. 2016;537:50–56. doi: 10.1038/nature19323. A human monoclonal antibody, aducanumab, reduces brain Aβ aggregates in AD patients in a dose- and time-dependent manner. This reduction in Aβ aggregates is accompanied by a slowing of cognitive decline, which will be of great interest to confirm in ongoing phase 3 clinical trials. [DOI] [PubMed] [Google Scholar]
  • 62.Solomon B, Koppel R, Frankel D, Hanan-Aharon E. Disaggregation of Alzheimer beta-amyloid by site-directed mAb. Proc Natl Acad Sci U S A. 1997;94:4109–4112. doi: 10.1073/pnas.94.8.4109. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 63.Mamikonyan G, Necula M, Mkrtichyan M, Ghochikyan A, Petrushina I, Movsesyan N, Mina E, Kiyatkin A, Glabe CG, Cribbs DH, et al. Anti-A beta 1–11 antibody binds to different beta-amyloid species, inhibits fibril formation, and disaggregates preformed fibrils but not the most toxic oligomers. J Biol Chem. 2007;282:22376–22386. doi: 10.1074/jbc.M700088200. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 64.Nguyen TX, Jaru-Ampornpan P, Lam VQ, Cao P, Piszkiewicz S, Hess S, Shan SO. Mechanism of an ATP-independent protein disaggregase: I. structure of a membrane protein aggregate reveals a mechanism of recognition by its chaperone. J Biol Chem. 2013;288:13420–13430. doi: 10.1074/jbc.M113.462812. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 65.Jaru-Ampornpan P, Shen K, Lam VQ, Ali M, Doniach S, Jia TZ, Shan SO. ATP-independent reversal of a membrane protein aggregate by a chloroplast SRP subunit. Nat Struct Mol Biol. 2010;17:696–702. doi: 10.1038/nsmb.1836. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 66.Jaru-Ampornpan P, Liang FC, Nisthal A, Nguyen TX, Wang P, Shen K, Mayo SL, Shan SO. Mechanism of an ATP-independent protein disaggregase: II. distinct molecular interactions drive multiple steps during aggregate disassembly. J Biol Chem. 2013;288:13431–13445. doi: 10.1074/jbc.M113.462861. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 67.Chakraborty A, Das I, Datta R, Sen B, Bhattacharyya D, Mandal C, Datta AK. A single-domain cyclophilin from Leishmania donovani reactivates soluble aggregates of adenosine kinase by isomerase-independent chaperone function. J Biol Chem. 2002;277:47451–47460. doi: 10.1074/jbc.M204827200. [DOI] [PubMed] [Google Scholar]
  • 68.Sen B, Chakraborty A, Datta R, Bhattacharyya D, Datta AK. Reversal of ADP-mediated aggregation of adenosine kinase by cyclophilin leads to its reactivation. Biochemistry. 2006;45:263–271. doi: 10.1021/bi0518489. [DOI] [PubMed] [Google Scholar]
  • 69.Mukherjee D, Patra H, Laskar A, Dasgupta A, Maiti NC, Datta AK. Cyclophilin-mediated reactivation pathway of inactive adenosine kinase aggregates. Arch Biochem Biophys. 2013;537:82–90. doi: 10.1016/j.abb.2013.06.018. [DOI] [PubMed] [Google Scholar]
  • ••70.Poepsel S, Sprengel A, Sacca B, Kaschani F, Kaiser M, Gatsogiannis C, Raunser S, Clausen T, Ehrmann M. Determinants of amyloid fibril degradation by the PDZ protease HTRA1. Nat Chem Biol. 2015;11:862–869. doi: 10.1038/nchembio.1931. In this study, human HtrA1 is revealed as a potent, ATP-independent protein disaggregase, which can dissolve and degrade tau and Aβ fibrils connected with AD. An engineered form of HtrA1 with low proteolytic activity dissolves tau and Aβ fibrils without degrading them. [DOI] [PubMed] [Google Scholar]
  • ••71.Krishnan R, Tsubery H, Proschitsky MY, Asp E, Lulu M, Gilead S, Gartner M, Waltho JP, Davis PJ, Hounslow AM, et al. A bacteriophage capsid protein provides a general amyloid interaction motif (GAIM) that binds and remodels misfolded protein assemblies. J Mol Biol. 2014;426:2500–2519. doi: 10.1016/j.jmb.2014.04.015. In this study, a bacteriophage capsid protein, gene 3 protein (g3p), is shown to enable amyloid-disaggregase activity of bacteriophages. A recombinant bivalent g3p molecule, an immunoglobulin Fc (Ig) fusion of the two N-terminal g3p domains (termed Ig-G3P), dissociated Aβ fibrils in an ATP-independent manner, and is now being developed as a potential therapeutic. [DOI] [PubMed] [Google Scholar]
  • 72.Riechmann L, Holliger P. The C-terminal domain of TolA is the coreceptor for filamentous phage infection of E. coli. Cell. 1997;90:351–360. doi: 10.1016/s0092-8674(00)80342-6. [DOI] [PubMed] [Google Scholar]
  • 73.Messing J. Phage M13 for the treatment of Alzheimer and Parkinson disease. Gene. 2016;583:85–89. doi: 10.1016/j.gene.2016.02.005. [DOI] [PubMed] [Google Scholar]
  • 74.Evans ML, Chapman MR. Curli biogenesis: order out of disorder. Biochim Biophys Acta. 2014;1843:1551–1558. doi: 10.1016/j.bbamcr.2013.09.010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • •75.Levenson JM, Schroeter S, Carroll JC, Cullen V, Asp E, Proschitsky M, Chung CH-Y, Gilead S, Nadeem M, Dodiya HB, et al. NPT088 reduces both amyloid-beta and tau pathologies in transgenic mice. Alzheimer’s & Dementia: Translational Research & Clinical Interventions. 2016;2:141–155. doi: 10.1016/j.trci.2016.06.004. Ig-G3P (renamed NPT088) rescues Aβ and tau pathologies in transgenic mice models of AD. NPT088 is now entering Phase 1b clinical trials for AD. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 76.Duennwald ML, Echeverria A, Shorter J. Small heat shock proteins potentiate amyloid dissolution by protein disaggregases from yeast and humans. PLoS Biol. 2012;10:e1001346. doi: 10.1371/journal.pbio.1001346. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 77.Gao X, Carroni M, Nussbaum-Krammer C, Mogk A, Nillegoda NB, Szlachcic A, Guilbride DL, Saibil HR, Mayer MP, Bukau B. Human Hsp70 Disaggregase Reverses Parkinson’s-Linked alpha-Synuclein Amyloid Fibrils. Mol Cell. 2015;59:781–793. doi: 10.1016/j.molcel.2015.07.012. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 78.Mattoo RU, Sharma SK, Priya S, Finka A, Goloubinoff P. Hsp110 is a bona fide chaperone using ATP to unfold stable misfolded polypeptides and reciprocally collaborate with Hsp70 to solubilize protein aggregates. J Biol Chem. 2013;288:21399–21411. doi: 10.1074/jbc.M113.479253. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 79.Nillegoda NB, Kirstein J, Szlachcic A, Berynskyy M, Stank A, Stengel F, Arnsburg K, Gao X, Scior A, Aebersold R, et al. Crucial HSP70 co-chaperone complex unlocks metazoan protein disaggregation. Nature. 2015;524:247–251. doi: 10.1038/nature14884. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 80.Rampelt H, Kirstein-Miles J, Nillegoda NB, Chi K, Scholz SR, Morimoto RI, Bukau B. Metazoan Hsp70 machines use Hsp110 to power protein disaggregation. EMBO J. 2012;31:4221–4235. doi: 10.1038/emboj.2012.264. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • ••81.Hjerpe R, Bett JS, Keuss MJ, Solovyova A, McWilliams TG, Johnson C, Sahu I, Varghese J, Wood N, Wightman M, et al. UBQLN2 Mediates Autophagy-Independent Protein Aggregate Clearance by the Proteasome. Cell. 2016;166:935–949. doi: 10.1016/j.cell.2016.07.001. In this study, Ubiquilin 2 is shown to act with the Hsp110, Hsp70, and Hsp40 disaggregase machinery to target disaggregated protein for degradation by the proteasome. Ubiquilin 2 recognizes client-bound Hsp70 and enables transfer of the client to the proteasome. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 82.Finka A, Sharma SK, Goloubinoff P. Multi-layered molecular mechanisms of polypeptide holding, unfolding and disaggregation by HSP70/HSP110 chaperones. Front Mol Biosci. 2015;2:29. doi: 10.3389/fmolb.2015.00029. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 83.Nillegoda NB, Bukau B. Metazoan Hsp70-based protein disaggregases: emergence and mechanisms. Front Mol Biosci. 2015;2:57. doi: 10.3389/fmolb.2015.00057. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 84.Hinault MP, Cuendet AF, Mattoo RU, Mensi M, Dietler G, Lashuel HA, Goloubinoff P. Stable alpha-synuclein oligomers strongly inhibit chaperone activity of the Hsp70 system by weak interactions with J-domain co-chaperones. J Biol Chem. 2010;285:38173–38182. doi: 10.1074/jbc.M110.127753. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • •85.Mashaghi A, Bezrukavnikov S, Minde DP, Wentink AS, Kityk R, Zachmann-Brand B, Mayer MP, Kramer G, Bukau B, Tans SJ. Alternative modes of client binding enable functional plasticity of Hsp70. Nature. 2016;539:448–451. doi: 10.1038/nature20137. In this study, alternative modes of client binding enable functional plasticity of Hsp70. The diversity of binding modes enables Hsp70 to stabilize and destabilize folded structures. [DOI] [PubMed] [Google Scholar]
  • 86.Eroglu B, Moskophidis D, Mivechi NF. Loss of Hsp110 leads to age-dependent tau hyperphosphorylation and early accumulation of insoluble amyloid beta. Mol Cell Biol. 2010;30:4626–4643. doi: 10.1128/MCB.01493-09. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 87.Cherkasov V, Hofmann S, Druffel-Augustin S, Mogk A, Tyedmers J, Stoecklin G, Bukau B. Coordination of translational control and protein homeostasis during severe heat stress. Curr Biol. 2013;23:2452–2462. doi: 10.1016/j.cub.2013.09.058. [DOI] [PubMed] [Google Scholar]
  • 88.Kroschwald S, Maharana S, Mateju D, Malinovska L, Nuske E, Poser I, Richter D, Alberti S. Promiscuous interactions and protein disaggregases determine the material state of stress-inducible RNP granules. Elife. 2015;4:e06807. doi: 10.7554/eLife.06807. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • ••89.Ganassi M, Mateju D, Bigi I, Mediani L, Poser I, Lee HO, Seguin SJ, Morelli FF, Vinet J, Leo G, et al. A Surveillance Function of the HSPB8-BAG3-HSP70 Chaperone Complex Ensures Stress Granule Integrity and Dynamism. Mol Cell. 2016;63:796–810. doi: 10.1016/j.molcel.2016.07.021. In this study, it is shown that only a minor fraction of aberrant stress granules is degraded by autophagy. The majority of stress granules are disassembled in a process that requires assistance by a HSPB8-BAG3-HSP70 chaperone complex. [DOI] [PubMed] [Google Scholar]
  • 90.Walters RW, Muhlrad D, Garcia J, Parker R. Differential effects of Ydj1 and Sis1 on Hsp70-mediated clearance of stress granules in Saccharomyces cerevisiae. RNA. 2015;21:1660–1671. doi: 10.1261/rna.053116.115. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 91.Kilpatrick K, Novoa JA, Hancock T, Guerriero CJ, Wipf P, Brodsky JL, Segatori L. Chemical induction of Hsp70 reduces alpha-synuclein aggregation in neuroglioma cells. ACS Chem Biol. 2013;8:1460–1468. doi: 10.1021/cb400017h. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 92.Kuo Y, Ren S, Lao U, Edgar BA, Wang T. Suppression of polyglutamine protein toxicity by co-expression of a heat-shock protein 40 and a heat-shock protein 110. Cell Death Dis. 2013;4:e833. doi: 10.1038/cddis.2013.351. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • ••93.Nagy M, Fenton WA, Li D, Furtak K, Horwich AL. Extended survival of misfolded G85R SOD1-linked ALS mice by transgenic expression of chaperone Hsp110. Proc Natl Acad Sci U S A. 2016;113:5424–5428. doi: 10.1073/pnas.1604885113. In this study, it is shown that neurodegeneration associated with cytosolic misfolding and aggregation of ALS-linked SOD1 can be ameliorated by overexpression of Hsp110, likely via stimulating the endogenous Hsp110, Hsp70, and Hsp40 disaggregase machinery. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 94.Mack KL, Shorter J. Engineering and Evolution of Molecular Chaperones and Protein Disaggregases with Enhanced Activity. Front Mol Biosci. 2016;3:8. doi: 10.3389/fmolb.2016.00008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 95.Aponte RA, Zimmermann S, Reinstein J. Directed evolution of the DnaK chaperone: mutations in the lid domain result in enhanced chaperone activity. J Mol Biol. 2010;399:154–167. doi: 10.1016/j.jmb.2010.03.060. [DOI] [PubMed] [Google Scholar]
  • 96.Aprile FA, Sormanni P, Vendruscolo M. A Rational Design Strategy for the Selective Activity Enhancement of a Molecular Chaperone toward a Target Substrate. Biochemistry. 2015;54:5103–5112. doi: 10.1021/acs.biochem.5b00459. [DOI] [PubMed] [Google Scholar]
  • 97.Needham PG, Masison DC. Prion-impairing mutations in Hsp70 chaperone Ssa1: effects on ATPase and chaperone activities. Arch Biochem Biophys. 2008;478:167–174. doi: 10.1016/j.abb.2008.07.023. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • •98.Fernandez-Funez P, Sanchez-Garcia J, de Mena L, Zhang Y, Levites Y, Khare S, Golde TE, Rincon-Limas DE. Holdase activity of secreted Hsp70 masks amyloid-beta42 neurotoxicity in Drosophila. Proc Natl Acad Sci U S A. 2016;113:E5212–5221. doi: 10.1073/pnas.1608045113. In this study, an engineered version of Hsp70 that gets secreted is shown to mitigate Aβ42 toxicity in Drosophila by promoting the formation of large, non-toxic Aβ42 assemblies. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 99.Mannini B, Cascella R, Zampagni M, van Waarde-Verhagen M, Meehan S, Roodveldt C, Campioni S, Boninsegna M, Penco A, Relini A, et al. Molecular mechanisms used by chaperones to reduce the toxicity of aberrant protein oligomers. Proc Natl Acad Sci U S A. 2012;109:12479–12484. doi: 10.1073/pnas.1117799109. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 100.Clausen T, Kaiser M, Huber R, Ehrmann M. HTRA proteases: regulated proteolysis in protein quality control. Nat Rev Mol Cell Biol. 2011;12:152–162. doi: 10.1038/nrm3065. [DOI] [PubMed] [Google Scholar]
  • 101.Tennstaedt A, Popsel S, Truebestein L, Hauske P, Brockmann A, Schmidt N, Irle I, Sacca B, Niemeyer CM, Brandt R, et al. Human high temperature requirement serine protease A1 (HTRA1) degrades tau protein aggregates. J Biol Chem. 2012;287:20931–20941. doi: 10.1074/jbc.M111.316232. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 102.Unal Gulsuner H, Gulsuner S, Mercan FN, Onat OE, Walsh T, Shahin H, Lee MK, Dogu O, Kansu T, Topaloglu H, et al. Mitochondrial serine protease HTRA2 p.G399S in a kindred with essential tremor and Parkinson disease. Proc Natl Acad Sci U S A. 2014;111:18285–18290. doi: 10.1073/pnas.1419581111. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 103.Ali YO, Li-Kroeger D, Bellen HJ, Zhai RG, Lu HC. NMNATs, evolutionarily conserved neuronal maintenance factors. Trends Neurosci. 2013;36:632–640. doi: 10.1016/j.tins.2013.07.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 104.Canto C, Menzies KJ, Auwerx J. NAD(+) Metabolism and the Control of Energy Homeostasis: A Balancing Act between Mitochondria and the Nucleus. Cell Metab. 2015;22:31–53. doi: 10.1016/j.cmet.2015.05.023. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 105.Zhai RG, Zhang F, Hiesinger PR, Cao Y, Haueter CM, Bellen HJ. NAD synthase NMNAT acts as a chaperone to protect against neurodegeneration. Nature. 2008;452:887–891. doi: 10.1038/nature06721. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 106.Ljungberg MC, Ali YO, Zhu J, Wu CS, Oka K, Zhai RG, Lu HC. CREB-activity and nmnat2 transcription are down-regulated prior to neurodegeneration, while NMNAT2 over-expression is neuroprotective, in a mouse model of human tauopathy. Hum Mol Genet. 2012;21:251–267. doi: 10.1093/hmg/ddr492. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • ••107.Ali YO, Allen HM, Yu L, Li-Kroeger D, Bakhshizadehmahmoudi D, Hatcher A, McCabe C, Xu J, Bjorklund N, Taglialatela G, et al. NMNAT2:HSP90 Complex Mediates Proteostasis in Proteinopathies. PLoS Biol. 2016;14:e1002472. doi: 10.1371/journal.pbio.1002472. In this study, NMNAT2 is shown to collaborate with Hsp90 to promote protein disaggregation and reactivation. NMNAT2 is proposed to function as a chaperone and disaggregase to reduce proteotoxic stress, while its biosynthetic activity protects neurons from excitotoxicity. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 108.Zhou M, Ottenberg G, Sferrazza GF, Hubbs C, Fallahi M, Rumbaugh G, Brantley AF, Lasmezas CI. Neuronal death induced by misfolded prion protein is due to NAD+ depletion and can be relieved in vitro and in vivo by NAD+ replenishment. Brain. 2015;138:992–1008. doi: 10.1093/brain/awv002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 109.Carlomagno Y, Zhang Y, Davis M, Lin WL, Cook C, Dunmore J, Tay W, Menkosky K, Cao X, Petrucelli L, et al. Casein kinase II induced polymerization of soluble TDP-43 into filaments is inhibited by heat shock proteins. PLoS One. 2014;9:e90452. doi: 10.1371/journal.pone.0090452. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 110.Jinwal UK, Akoury E, Abisambra JF, O’Leary JC, 3rd, Thompson AD, Blair LJ, Jin Y, Bacon J, Nordhues BA, Cockman M, et al. Imbalance of Hsp70 family variants fosters tau accumulation. FASEB J. 2013;27:1450–1459. doi: 10.1096/fj.12-220889. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 111.Lek M, Karczewski KJ, Minikel EV, Samocha KE, Banks E, Fennell T, O’Donnell-Luria AH, Ware JS, Hill AJ, Cummings BB, et al. Analysis of protein-coding genetic variation in 60,706 humans. Nature. 2016;536:285–291. doi: 10.1038/nature19057. [DOI] [PMC free article] [PubMed] [Google Scholar]

RESOURCES