Next Article in Journal
Human Cytochrome P450 1, 2, 3 Families as Pharmacogenes with Emphases on Their Antimalarial and Antituberculosis Drugs and Prevalent African Alleles
Previous Article in Journal
Cost-Effective Next Generation Sequencing-Based STR Typing with Improved Analysis of Minor, Degraded and Inhibitor-Containing DNA Samples
Previous Article in Special Issue
Intrinsically Disordered Proteins: An Overview
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Aggregation of Disordered Proteins Associated with Neurodegeneration

by
Phoebe S. Tsoi
,
My Diem Quan
,
Josephine C. Ferreon
* and
Allan Chris M. Ferreon
*
Department of Pharmacology and Chemical Biology, Baylor College of Medicine, Houston, TX 77030, USA
*
Authors to whom correspondence should be addressed.
Int. J. Mol. Sci. 2023, 24(4), 3380; https://doi.org/10.3390/ijms24043380
Submission received: 14 January 2023 / Revised: 2 February 2023 / Accepted: 6 February 2023 / Published: 8 February 2023
(This article belongs to the Special Issue Intrinsically Disordered Proteins (IDPs) 2.0)

Abstract

:
Cellular deposition of protein aggregates, one of the hallmarks of neurodegeneration, disrupts cellular functions and leads to neuronal death. Mutations, posttranslational modifications, and truncations are common molecular underpinnings in the formation of aberrant protein conformations that seed aggregation. The major proteins involved in neurodegeneration include amyloid beta (Aβ) and tau in Alzheimer’s disease, α-synuclein in Parkinson’s disease, and TAR DNA-binding protein (TDP-43) in amyotrophic lateral sclerosis (ALS). These proteins are described as intrinsically disordered and possess enhanced ability to partition into biomolecular condensates. In this review, we discuss the role of protein misfolding and aggregation in neurodegenerative diseases, specifically highlighting implications of changes to the primary/secondary (mutations, posttranslational modifications, and truncations) and the quaternary/supramolecular (oligomerization and condensation) structural landscapes for the four aforementioned proteins. Understanding these aggregation mechanisms provides insights into neurodegenerative diseases and their common underlying molecular pathology.

1. Introduction

Neurodegenerative diseases (NDs) are characterized by the progressive damage and dysfunction of neuronal cells. Despite exhibiting diverse clinical symptoms, these diseases share many common pathologic features. ND pathogenesis is thought to involve protein misfolding and aggregation within specific brain regions, which trigger neuroinflammation and oxidative stress at sites of injury, subsequently leading to degeneration of neural tissues. Molecular mechanisms responsible for the initial protein misfolding and the change from functional protein to pathologic aggregates have been a major subject of research in recent years.
NDs are classified according to the clinical symptoms and the major component(s) of protein deposits found in affected brain regions [1]. A vast majority of these diseases are associated with a class of proteins known as intrinsically disordered proteins, or IDPs. These proteins can be entirely unstructured or hybrids of structured domains and long stretches of intrinsically disordered regions (IDRs). IDRs are primarily composed of polar or charged amino acids, lacking sufficient hydrophobic residues that can mediate cooperative folding [2]. IDPs exist in a dynamic equilibrium of multiple conformational states of varying degrees of folding under physiological conditions [3,4,5]. The conformations adopted are largely affected by factors such as amino acid sequence, embedded motifs, and charge distribution/arrangement [6,7,8].
IDPs are highly prevalent in many proteomes, including that of humans, and play important roles in cellular processes such as the regulation of transcription and translation [9,10], cell cycle control [11,12], and cell signaling [12,13]. Changes in the cellular milieu and/or mutation(s) in IDPs can disrupt normal protein functions, resulting in misfolding and aggregation/fibrillation [14,15]. Misfolded proteins can serve as conformational switches and/or seeds that proceed to self-propagation [16,17,18,19], taking on prion-like properties and causing cellular stress and damage [20,21]. Misfolded proteins can also cross-seed and induce other proteins to aggregate [22,23,24,25]. IDPs associated with common neurodegenerative diseases include amyloid beta (Aβ) and tau for Alzheimer’s disease (AD), α-synuclein (α-syn) for Parkinson’s disease (PD), and TAR DNA-binding protein (TDP-43) for amyotrophic lateral sclerosis (ALS) (Figure 1).
Aβ is a small, 39–43 amino acid peptide heavily implicated in AD. The peptide was identified to be the main component of neocortical plaques [34], a pathological hallmark of AD [35]. AlphaFold predicts Aβ-42 to be a primarily disordered peptide with a small helical region, whereas IUPred assigned a low disorder score. Experimentally, Aβ-42 exists as a mixture of random coils, α-helixes, and β-sheets, and specific conformations are favored based on environmental conditions [36,37], suggesting that Aβ-42 can be induced to favor conformations that are more aggregation prone. Tau is a microtubule-associated protein that exists as various isoforms with varying numbers of microtubule binding domains and 29-amino-acid-long inserts [38]. The isoform shown in Figure 1 contains three microtubule-binding repeats (tau-3R) and exhibits high disorder according to both AlphaFold and IUPred predictions. Tau is a major component of neurofibrillary tangles (NFTs) in AD [39]. α-Syn is a 140 amino acid protein whose mutated forms were first discovered in relation to PD [40]. α-Syn encodes a helical N-terminus, a non-amyloidal component domain (NAC), and an acidic, disordered C-terminal domain. TDP-43 is a 414 amino acid protein implicated in over 90% of ALS cases [41,42]. It contains a structured N-terminal domain, two RNA recognition motifs (RRMs), and a disordered low complexity C-terminal domain (LCD).
The four aforementioned proteins undergo misfolding from their native states to form β-sheet-rich structures ranging from small oligomers to large fibrillar aggregates in diseased brains [43,44]. This review describes the molecular precursors responsible for misfolding and aggregation of the aforementioned ND-associated IDPs, summarized in Figure 2.

2. Mutation

AD, PD, and ALS occur in both sporadic and familial forms. The majority of ND cases are sporadic, constituting approximately 97% of AD cases [45], 85% of PD cases [46], and 90–95% of ALS cases [47]. The remaining familial occurrences are caused by inheritable mutations, which in disease-related proteins generally exhibit similar phenotypes: increased aggregation propensities [48,49,50], altered protein populations [51,52,53] and proteasomal regulation [51,54], and increased cytotoxicity in model organisms [55,56,57].
Many Aβ mutations are located on the APP gene, which encodes for the amyloid precursor protein (APP) [58]. Most of the pathogenic mutations occur in a section of the APP gene that encodes the proteolytic sites of β- and γ-secretases, which result in an overall increased generation of Aβ [54,59]. Other APP mutations can affect the cleavage process from APP to Aβ and generate truncated products exhibiting differential aggregation propensities [52]. In particular, the Arctic APP mutant (APP E693G) displays high levels of protofibrils as well as cognitive defects in mice [55,60,61].
Tau protein is encoded by the MAPT gene, with both exonic and intronic mutations identified in tauopathies. Mutations generally promote tau aggregation by altering the ratio of tau containing three tubulin-binding repeats (tau-3R) to four tubulin-binding repeats (tau-4R), otherwise known as the 3R:4R tau ratio [48,49]. Tau-3R and tau-4R isoforms exist in a one-to-one ratio in most regions of the brain; deviations from this ratio characterize tauopathies [38]. Increased tau-4R expression has been found to promote tau phosphorylation and oligomerization and induce behavioral abnormalities in a mouse model expressing human tau [56]. Mutations also induce tau fragmentation and enhance tau hyperphosphorylation [51,62]. Extensively studied tau mutants include G272V, P301L, V337M, and R406W [63,64,65,66]; however, it should be noted that these mutants are associated with frontotemporal dementia (FTD), and not AD. No MAPT mutations have been associated with AD so far, suggesting that mechanisms underlying tau aggregation in AD may be different from those involved in other tauopathies caused by MAPT mutations.
α-Syn was first implicated in NDs when the A53T mutation was identified in autosomal dominant PD [40]. Autosomal dominant forms of PD are associated with N-terminal missense mutations, such as A53T, A53E, A30P, and E46K. [67,68,69]. Mutants A53T and A30P of α-syn have been shown to be structurally defective for membrane binding [31,32,70,71] and exhibit enhanced self-aggregation propensity and kinetics [72]. Mutant A30P α-syn has also been shown to exhibit 2-state folding thermodynamics, compared to the 3-state folding behavior of wild-type α-syn [71]. Wild-type α-syn is capable of assembling into two types of dimers, with one dimeric form more favored than the other [73]. Mutants A53T, A30P, and E46K have been demonstrated to promote dimerization and enhance the formation of the less favored dimeric structure [73]. The structural heterogeneity of α-syn dimers is suggested to indicate different aggregation pathways. More recently discovered mutations, A18T and A29S, were found to aggregate faster than wild-type α-syn, with the A18T mutant having faster aggregation kinetics compared to A29S [74].
Numerous mutations in the TARDBP gene have been identified as being associated with ALS. These TDP-43 mutations can increase aggregation propensity, enhance cytoplasmic mislocalization, and alter protein stability [50,75]. Most ALS-associated mutations appear in exon 6 of the TARDBP gene, which encodes for the intrinsically disordered C-terminal region of TDP-43. The most well-studied TDP-43 mutations include A315T, Q331K, M337V, and D169G, for which several ALS disease models have also been established [50]. Recombinantly expressed TDP-43 containing ALS-linked mutations were found to have increased aggregation in vitro and promoted cytotoxicity in yeast cells [57]. Peptides from the TDP-43 amyloidogenic core region (residues 286–366) containing ALS-associated mutations also form amyloid-like fibrils [76,77]. TDP-43 A315T mutant has been found to form amyloid fibrils in vitro and cause cell death when added to cultured neuronal cells [77]. Additional information regarding mutations, as well as reviews that discuss the role of mutations in NDs in-depth, can be found in Table 1.

3. Posttranslational Modification

Posttranslational modifications (PTMs) are frequently used to regulate IDP function, localization, and turnover [85,86]. Aberrant PTMs disrupt Aβ, tau, α-syn, and TDP-43 functions and are linked to neurodegeneration. The mechanisms by which PTM leads to disease pathology are largely dependent on the type of PTM and the protein involved. In general, however, PTMs, such as phosphorylation and acetylation, alter charge properties, affect binding interactions [87], folding and conformational stability [88], and oligomerization states [89], all of which exhibit the ability to modulate aggregation. Pathologic PTMs include phosphorylation, ubiquitination, acetylation, and glycosylation.
PTMs of Aβ have been demonstrated to increase its aggregation rate. Phosphorylation of residue S8 increases the stability of the β-sheet conformation of Aβ [90,91], and phosphorylated Aβ have been detected in the brains of transgenic mice and AD patients [88]. In Drosophila, phosphorylated Aβ induced higher toxicity compared to non-phosphorylated Aβ [90]. In addition to phosphorylation, glycation also stimulates amyloid aggregation. Aβ in amyloid deposits are glycated [92]; advanced glycation of Aβ can seed and accelerate aggregation of soluble Aβ peptide [93,94]. Cerebral spinal fluid of AD patients contain abnormally O-glycosylated Aβ peptides of 15–17 residue size [95], suggesting that glycosylation may play a role in Aβ clearance [96].
Tau hyperphosphorylation is responsible for its loss of physiological functions, gain in toxicity, and aggregation in the form of NFTs. Specifically, residues S396, S404, and S422 have been found abnormally hyperphosphorylated in diseased brains [97], and it has been suggested that phosphorylation at these residues may influence tau aggregation [98]. Hyperphosphorylation has also been shown to impair microtubule binding [87], which could lead to microtubule destabilization and compromised cytoskeletal integrity. Tau acetylation was first recognized in ND mouse models [99,100]. Acetylation of K280 weakens the binding of tau to negatively charged microtubules, potentially destabilizing microtubule networks [101,102]. Ubiquitin has been identified in tau inclusions extracted from the brains of tauopathy patients; however, the role of ubiquitination in ND is not yet known. While ubiquitin is a component of tau aggregates found in the brains of AD patients, tau phosphorylation precedes its ubiquitination in the NFTs of AD patients [103,104]. This suggests that ubiquitin may be linked to tau after the formation of the fibrillar inclusions. However, other groups have reported that both mono- and polyubiquitination contribute to the formation of insoluble protein inclusions present in neurodegenerative diseases [105,106], and that tau ubiquitination in cell cultures increases aggregation [107].
Addition of side chain modifiers has also been shown to modulate α-syn toxicity. Physiological levels of phosphorylated α-syn are relatively low; however, threonine, serine, and tyrosine hyperphosphorylation are commonly found in pathologically aggregated α-syn [108,109]. Most phosphorylated residues are located in the C-terminus of α-syn, which is thought to be involved in α-syn pathology. S129, in particular, is phosphorylated in >90% of PD patients and is used as a pathological marker [108]. Ubiquitinated α-syn is often present in inclusions of PD patients in conjunction with phosphorylation [110]. Sumoylation [111,112], nitration [113], and glycosylation [114] of α-syn have also been observed in association with α-syn toxicity.
The two most pathologically significant PTMs in TDP-43 are phosphorylation and ubiquitination. TDP-43 phosphorylation is a signature of ALS pathology; S409/S410 phosphorylation, in particular, are distinctly observed in ALS patients [115]. Phosphorylation is associated with cytoplasmic mislocalization and aggregation of TDP-43 in neurons [116,117]. TDP-43 has also been found in the ubiquitinated state in ALS brain inclusions [42]. Ubiquitination facilitates TDP-43 cytoplasmic accumulation into inclusions without any detectable evidence of its degradation [118,119]. Additionally, acetylation may serve a pivotal role in mediating TDP-43 function and dysfunction [120].
PTM is an ever-expanding field of research. Although the PTMs most discussed in-depth in this review were phosphorylation and ubiquitination, many other PTMs affect these proteins and other disease-related IDPs. Table 2 provides references to additional resources that discuss the roles of PTM addition for each protein. Further investigation of other covalent modifications is warranted to build a more comprehensive picture of how the cells regulate behaviors of ND-associated proteins and how aberrant regulations can be rescued.

4. Truncation

Protein truncation is one of the most common pathological modifications of IDPs. Truncations in IDPs can alleviate steric hindrance involved with protein folding, inducing structural changes that can lead to protein misfolding [126]. The truncated forms exhibiting perturbed aggregation behavior can act as seeds for nucleation and partake in self-assembly that results in the formation of insoluble structures. Moreover, truncation is known to impede vital functions of IDPs, leading to both loss-of-function and toxic gain-of-function. A plethora of studies have demonstrated that truncated forms of neuronal IDPs are the driving force in various neurodegenerative proteinopathies [51,127,128].
Aβ peptides are generated by the cleavage of APP via β- and γ-secretases [129]. In nonneuronal cells, the majority of APP proteins are initially cleaved by α-secretase. This nonamyloidogenic cleavage occurs within the Aβ domain and prevents the production of Aβ [130]. In contrast, through the amyloidogenic pathway, β-secretase mediates the initial cleavage of APP, which is subsequently processed by γ-secretase to produce Aβ peptides [131,132]. The amyloidogenic cleavage occurs within the transmembrane domain of APP and generates C-terminally truncated peptides of various sizes, ranging from 38 to 42 amino acids. The levels of generated peptides are used to distinguish AD from other NDs [133,134]. The Aβ species most strongly implicated in AD is Aβ-42. Aβ-42 exhibits enhanced aggregation propensity compared to other Aβ peptides [135]. Although healthy individuals can also generate Aβ peptides, higher levels of Aβ-42 are detected in AD patients’ brain samples [133,136], and AD patients have been observed to generate longer Aβ forms compared to unafflicted individuals [127].
Truncated tau species are derived from proteolytic processing via proteases, of which caspases and calpains are of particular interest. Specific cleavage products of caspases −2, −3, and −6 have been linked to AD. A truncated form of tau generated by caspase-2 cleavage at D314 has been found in AD brains [51,137]. Caspase-3 cleavage at D421 generates a tau-421 species; elevated levels of caspase-3 and tau-421 have been observed in AD [51,138]. Tau-421 colocalizes with NFTs in human AD brain and correlates with NFT formation and cognitive impairment in aged mice [139]. Tau can also be cleaved by caspase-6 to produce tau-13 and tau-402 truncations. Active caspase-6 and tau-402 were observed in NFTs and neuritic plaques in the AD brain [140]. In addition, tau-402 levels in cerebrospinal fluid correlate with impaired cognitive performance in AD patients [141]. Calpains are calcium-dependent cysteine proteases, and calpain-mediated tau cleavage generates several truncated tau isoforms such as tau-45–230 and tau-243–441 [142,143]. Increased levels of tau-45–230 have been identified in AD brain samples, and elevated tau-243–441 levels are observed in transgenic tau Tg601 mouse model [143].
α-Syn inclusions in human brain contain C-terminally truncated α-syn protein, which may be generated by proteasome degradation or calpain cleavage [144,145]. C-terminally truncated α-syn fibrillizes in vitro [146], and mice expressing this α-syn species in dopaminergic neurons demonstrate neuronal aggregates with either granular or fibrillar morphologies [147]. Caspase cleavage of α-syn generates 1–121 α-syn. This truncated peptide assembles into fibrils and demonstrates prion-like seeding [128,148].
The C-terminal region of TDP-43 is highly disordered and comprises a glycine-rich region. Highly cytotoxic C-terminal fragments of sizes 25–35 kDa that are produced from aberrant caspase cleavage of TDP-43 are the prominent species found in the inclusion bodies identified from ALS-affected brains [149,150]. The C-terminal region of TDP-43 also contains a short, highly dynamic and unstable helix–turn–helix region in residues 311–360. Peptides containing this region form amyloid-like fibrils in vitro, which can exhibit prion-like toxicity in cells [76,151]. Table 3 provides additional resources that discuss different protein truncations and their roles in NDs.

5. Toxic Oligomerization

Significant evidence suggests that smaller, soluble misfolded oligomers may be the true cause of neurodegeneration [155,156,157]. Misfolded oligomers are a group of species that exist in a range of sizes, from dimers to protofibrillar structures [158,159,160]. These oligomeric species are highly dynamic and exist in equilibrium with monomers and fibrils. Some oligomers may be intermediates for amyloid fibril formation while others might be terminal, off-pathway products [160,161]. The heterogeneity, thermodynamic interconversion between species, and aggregation propensity of these oligomers have made it very difficult to obtain high-resolution structural information, as well as to determine which are the most relevant oligomeric structures for disease [157,159].
The Aβ oligomer hypothesis was introduced in 1998 and suggested that the damage found in AD patient brains is caused by oligomeric species [162]. Since then, oligomeric species of various sizes have been identified in human brains and in brains from APP transgenic mice, although the characterization of these oligomers has been hindered by their metastability and heterogeneity [159]. Oligomeric species that are ‘on-pathway’ and ‘off-pathway’ to fibril formation have been observed [163]. Some oligomers (>50 kDa) identified by mass spectrometry are toxic and off-pathway of amyloid formation [164,165] while other, smaller species (<50 kDa) that react with anti-fibril antibodies readily form fibrils and are less associated with toxicity [166,167]. Aβ can also assemble into intermediate structures known as protofibrils, which are large (>100 kDa) on-pathway species for fibrillation [168]. These Aβ protofibrillar species, which have been shown to induce neurotoxicity in rat cortical neurons [168] and impair cognitive and behavioral functions (such as spatial–temporal pattern separation and learning in mice [169]), are associated with inflammatory responses, and have been detected in AD brains [170].
Oligomeric tau may also be the toxic species in tauopathies. Hyperphosphorylated tau assembles into oligomers prior to NFT formation. Hyperphosphorylated tau monomers have lowered affinity to microtubules and increased affinity for other tau monomers to form oligomeric tau that is detergent-soluble. These tau oligomers potentiate neuronal damage, leading to neurodegeneration and traumatic brain injury [89,171,172]. In AD brain samples, tau oligomers were found at a fourfold greater concentration compared to healthy control samples [173]. When the oligomer grows, it adapts a β-sheet structure and transforms into a detergent-insoluble aggregate with granular appearance, which elongates into tau fibrils and ultimately forms NFTs [174]. This process suggests that tau oligomers may be involved in neuronal dysfunction prior to NFT formation [175].
α-Syn matures from monomer to amyloid fibrils rich in β-sheets through several intermediate oligomeric species [176,177]. Single-molecule studies of α-syn have identified conformations that can initiate pathologic aggregation [31,71,178]. Findings have suggested that certain α-syn oligomers or protofibrils may be toxic [179,180]. The initial observation that the mutant A30P α-syn monomers were consumed more rapidly but fibrillized more slowly than WT α-syn suggests that the oligomeric intermediary species may be pathologic, rather than the fibrillar forms [176,177]. Direct in vivo data supporting the toxic protofibril hypothesis are still relatively limited, and most of the evidence is circumstantial.
TDP-43 oligomerization and its potential neurotoxic properties have also been studied. In normal brains, TDP-43 exist as dimers in the nucleus of neurons [181,182]. However, there are reports of pathologic TDP-43 oligomers, which may be structurally distinct from the nuclear oligomers. The N-terminal domain regulates dimerization [181,183], and the N-terminal region (residues 3–183) acts as an intermolecular interacting domain for an 86 kDa species that was observed in an immunoblot of extracts from deceased ALS patient brains [184]. Full-length TDP-43, not only certain domains, has also been shown to form stable, spherical cytotoxic oligomers in neuronal cells [185]. These TDP-43 oligomers can also cross-seed Aβ-42 peptide, demonstrating a structural conversion that can occur among common amyloid species [185,186]. Furthermore, such TDP-43 oligomeric aggregates were detected in brain sections of TDP-43 mouse models as well as ND patients [185].

6. Biomolecular Condensation

Biomolecular condensates (BMCs) include membraneless organelles, which are small compartments that are not encased in lipid membranes. BMC formation can be driven by liquid–liquid phase separation (LLPS)—a process by which a solution demixes into two liquid states: one phase concentrated in macromolecules and another dilute phase. [187,188,189]. It should be noted that while IDPs typically show increased tendency to phase separate, disorder is not a prerequisite for phase separation. Structured domains can also be induced to phase separate [190]. BMCs congregate and concentrate proteins and nucleic acids that are involved in diverse processes, including RNA metabolism, DNA damage response, and signal transduction. Although these condensates play many physiological roles [191,192], their aberrant behaviors may be associated with disease, especially neurodegeneration [193,194,195].
In the case of Aβ, recruitment into BMCs may actually be protective. A recent study by Kuffner et al. demonstrated that Aβ-42 was recruited into a liquid droplet along with other scaffold proteins of membraneless organelles. Although Aβ-42 concentration increased in the sequestered phase, its aggregation was inhibited [196]. This result suggests that another role of BMCs is to sequester aggregation-prone proteins as a protective measure.
Tau phase separation plays a role in its biological and pathological functions. Tau decreases the critical concentration for tubulin assembly [197]. In the presence of tau condensates, tubulin assembly is further facilitated, indicating that tau condensates may act as nucleation sites for microtubules [191]. Phase separation may also initiate tau aggregation, as phosphorylated tau has higher phase separation propensity [193,198]. Tau phase separation has also been shown to be inhibited by acetylation [199]. These studies suggest that tau BMCs serve a functional role that can be modulated by PTMs, although further studies are required to fully elucidate the regulation of BMCs.
α-Syn has been shown to phase separate in vitro in the presence of molecular crowding agents, although at high critical concentrations and after prolonged incubation [200]. Mutations and PTMs significantly lower this critical concentration and decrease the time necessary to form droplets [194]. These observations suggest that the phase separation properties of α-syn are more closely related to its aggregation.
TDP-43 BMCs have both functional and dysfunctional relevance. TDP-43 is a component of stress granules, transient repositories of proteins and RNAs that form when cells are exposed to internal or external stressors [192,195]. TDP-43 is also found in insoluble cytoplasmic inclusions that are hypothesized to originate from stress granules [195]. In vitro studies of TDP-43 have shown that phase separation facilitates aggregation of the unstructured C-terminus [201], and that fibrillation and phase separation can be decoupled [202]. ALS-relevant TDP-43 mutants disrupt phase separation and enhance aggregation, though with irregular morphology [203].

7. Conclusions

With the emergence of IDP roles in vital cell functions, as well as their implications in various diseases, the field of IDP research has been rapidly expanding within recent decades. This review sought to summarize the molecular mechanisms that underlie the major IDPs associated with the most common neurodegenerative diseases. Mutations, aberrant PTMs, and truncations contribute to the pathologic properties of ND-associated proteins Aβ and tau, α-syn, and TDP-43, respectively, associated with AD, PD, and ALS. The capability of each protein to assemble into cytotoxic oligomeric species that may evolve into the aggregated hallmarks of their respective diseases as well as the functional and dysfunctional duality of their biomolecular condensates demonstrate that the pathway from functional to pathological is not straightforward, rather winding with multiple avenues for regulation and therapy. While this review is primarily focused on four specific proteins, these trends can be observed in other IDPs associated with neurodegenerative diseases such as TIA-1, hnRNP, FUS, ataxin, and SOD1. Continued research in this field will allow us to further understand the pathology of neurodegenerative diseases and develop effective therapeutic strategies.

Author Contributions

Writing—original draft preparation, P.S.T. and M.D.Q.; writing—review and editing, P.S.T., M.D.Q., J.C.F. and A.C.M.F. All authors have read and agreed to the published version of the manuscript.

Funding

Funding was provided by the National Institute of Neurological Disorders and Stroke (NINDS), National Institutes of Health (NIH) grants R01 NS105874, R21 NS107792, and R21 NS109678 to A.C.M.F. and by the National Institute of General Medical Sciences (NIGMS), NIH grant R01 GM122763 to J.C.F.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Kovacs, G.G. Molecular Pathological Classification of Neurodegenerative Diseases: Turning towards Precision Medicine. Int. J. Mol. Sci. 2016, 17, 189. [Google Scholar] [CrossRef] [PubMed]
  2. Uversky, V.N.; Gillespie, J.R.; Fink, A.L. Why are “natively unfolded” proteins unstructured under physiologic conditions? Proteins 2000, 41, 415–427. [Google Scholar] [CrossRef] [PubMed]
  3. Forman-Kay, J.D.; Mittag, T. From sequence and forces to structure, function, and evolution of intrinsically disordered proteins. Structure 2013, 21, 1492–1499. [Google Scholar] [CrossRef] [PubMed]
  4. Dyson, H.J.; Wright, P.E. Intrinsically unstructured proteins and their functions. Nat. Rev. Mol. Cell Biol. 2005, 6, 197–208. [Google Scholar] [CrossRef] [PubMed]
  5. Wei, G.; Xi, W.; Nussinov, R.; Ma, B. Protein Ensembles: How Does Nature Harness Thermodynamic Fluctuations for Life? The Diverse Functional Roles of Conformational Ensembles in the Cell. Chem. Rev. 2016, 116, 6516–6551. [Google Scholar] [CrossRef]
  6. Marsh, J.A.; Forman-Kay, J.D. Sequence determinants of compaction in intrinsically disordered proteins. Biophys. J. 2010, 98, 2383–2390. [Google Scholar] [CrossRef]
  7. Mao, A.H.; Crick, S.L.; Vitalis, A.; Chicoine, C.L.; Pappu, R.V. Net charge per residue modulates conformational ensembles of intrinsically disordered proteins. Proc. Natl. Acad. Sci. USA 2010, 107, 8183–8188. [Google Scholar] [CrossRef]
  8. Muller-Spath, S.; Soranno, A.; Hirschfeld, V.; Hofmann, H.; Ruegger, S.; Reymond, L.; Nettels, D.; Schuler, B. From the Cover: Charge interactions can dominate the dimensions of intrinsically disordered proteins. Proc. Natl. Acad. Sci. USA 2010, 107, 14609–14614. [Google Scholar] [CrossRef]
  9. Shammas, S.L. Mechanistic roles of protein disorder within transcription. Curr. Opin. Struct. Biol. 2017, 42, 155–161. [Google Scholar] [CrossRef]
  10. Tsafou, K.; Tiwari, P.B.; Forman-Kay, J.D.; Metallo, S.J.; Toretsky, J.A. Targeting Intrinsically Disordered Transcription Factors: Changing the Paradigm. J. Mol. Biol. 2018, 430, 2321–2341. [Google Scholar] [CrossRef]
  11. Yoon, M.K.; Mitrea, D.M.; Ou, L.; Kriwacki, R.W. Cell cycle regulation by the intrinsically disordered proteins p21 and p27. Biochem. Soc. Trans. 2012, 40, 981–988. [Google Scholar] [CrossRef]
  12. Bondos, S.E.; Dunker, A.K.; Uversky, V.N. On the roles of intrinsically disordered proteins and regions in cell communication and signaling. Cell Commun. Signal. 2021, 19, 88. [Google Scholar] [CrossRef]
  13. Wright, P.E.; Dyson, H.J. Intrinsically disordered proteins in cellular signalling and regulation. Nat. Rev. Mol. Cell Biol. 2015, 16, 18–29. [Google Scholar] [CrossRef]
  14. Oldfield, C.J.; Dunker, A.K. Intrinsically disordered proteins and intrinsically disordered protein regions. Annu. Rev. Biochem. 2014, 83, 553–584. [Google Scholar] [CrossRef]
  15. Radivojac, P.; Iakoucheva, L.M.; Oldfield, C.J.; Obradovic, Z.; Uversky, V.N.; Dunker, A.K. Intrinsic disorder and functional proteomics. Biophys. J. 2007, 92, 1439–1456. [Google Scholar] [CrossRef]
  16. Stefani, M.; Dobson, C.M. Protein aggregation and aggregate toxicity: New insights into protein folding, misfolding diseases and biological evolution. J. Mol. Med. 2003, 81, 678–699. [Google Scholar] [CrossRef]
  17. Breydo, L.; Redington, J.M.; Uversky, V.N. Effects of Intrinsic and Extrinsic Factors on Aggregation of Physiologically Important Intrinsically Disordered Proteins. Int. Rev. Cell Mol. Biol. 2017, 329, 145–185. [Google Scholar] [CrossRef]
  18. Candelise, N.; Scaricamazza, S.; Salvatori, I.; Ferri, A.; Valle, C.; Manganelli, V.; Garofalo, T.; Sorice, M.; Misasi, R. Protein Aggregation Landscape in Neurodegenerative Diseases: Clinical Relevance and Future Applications. Int. J. Mol. Sci. 2021, 22, 6016. [Google Scholar] [CrossRef]
  19. Soto, C.; Pritzkow, S. Protein misfolding, aggregation, and conformational strains in neurodegenerative diseases. Nat. Neurosci. 2018, 21, 1332–1340. [Google Scholar] [CrossRef]
  20. Hromadkova, L.; Siddiqi, M.K.; Liu, H.; Safar, J.G. Populations of Tau Conformers Drive Prion-like Strain Effects in Alzheimer’s Disease and Related Dementias. Cells 2022, 11, 2997. [Google Scholar] [CrossRef]
  21. Haynes, C.M.; Titus, E.A.; Cooper, A.A. Degradation of misfolded proteins prevents ER-derived oxidative stress and cell death. Mol. Cell 2004, 15, 767–776. [Google Scholar] [CrossRef] [PubMed]
  22. Furukawa, Y.; Kaneko, K.; Matsumoto, G.; Kurosawa, M.; Nukina, N. Cross-seeding fibrillation of Q/N-rich proteins offers new pathomechanism of polyglutamine diseases. J. Neurosci. 2009, 29, 5153–5162. [Google Scholar] [CrossRef] [PubMed]
  23. Lim, K.H. Diverse Misfolded Conformational Strains and Cross-seeding of Misfolded Proteins Implicated in Neurodegenerative Diseases. Front. Mol. Neurosci. 2019, 12, 158. [Google Scholar] [CrossRef] [PubMed]
  24. Subedi, S.; Sasidharan, S.; Nag, N.; Saudagar, P.; Tripathi, T. Amyloid Cross-Seeding: Mechanism, Implication, and Inhibition. Molecules 2022, 27, 1776. [Google Scholar] [CrossRef] [PubMed]
  25. Morales, R.; Moreno-Gonzalez, I.; Soto, C. Cross-seeding of misfolded proteins: Implications for etiology and pathogenesis of protein misfolding diseases. PLoS Pathog. 2013, 9, e1003537. [Google Scholar] [CrossRef]
  26. Jumper, J.; Evans, R.; Pritzel, A.; Green, T.; Figurnov, M.; Ronneberger, O.; Tunyasuvunakool, K.; Bates, R.; Zidek, A.; Potapenko, A.; et al. Highly accurate protein structure prediction with AlphaFold. Nature 2021, 596, 583–589. [Google Scholar] [CrossRef]
  27. Dosztanyi, Z. Prediction of protein disorder based on IUPred. Protein Sci. 2018, 27, 331–340. [Google Scholar] [CrossRef]
  28. Dosztanyi, Z.; Csizmok, V.; Tompa, P.; Simon, I. The pairwise energy content estimated from amino acid composition discriminates between folded and intrinsically unstructured proteins. J. Mol. Biol. 2005, 347, 827–839. [Google Scholar] [CrossRef]
  29. Meszaros, B.; Erdos, G.; Dosztanyi, Z. IUPred2A: Context-dependent prediction of protein disorder as a function of redox state and protein binding. Nucleic Acids Res. 2018, 46, W329–W337. [Google Scholar] [CrossRef]
  30. Meszaros, B.; Simon, I.; Dosztanyi, Z. Prediction of protein binding regions in disordered proteins. PLoS Comput. Biol. 2009, 5, e1000376. [Google Scholar] [CrossRef] [Green Version]
  31. Ferreon, A.C.; Gambin, Y.; Lemke, E.A.; Deniz, A.A. Interplay of alpha-synuclein binding and conformational switching probed by single-molecule fluorescence. Proc. Natl. Acad. Sci. USA 2009, 106, 5645–5650. [Google Scholar] [CrossRef]
  32. Ferreon, A.C.; Deniz, A.A. Alpha-synuclein multistate folding thermodynamics: Implications for protein misfolding and aggregation. Biochemistry 2007, 46, 4499–4509. [Google Scholar] [CrossRef]
  33. Trexler, A.J.; Rhoades, E. Alpha-synuclein binds large unilamellar vesicles as an extended helix. Biochemistry 2009, 48, 2304–2306. [Google Scholar] [CrossRef]
  34. Cras, P.; van Harskamp, F.; Hendriks, L.; Ceuterick, C.; van Duijn, C.M.; Stefanko, S.Z.; Hofman, A.; Kros, J.M.; Van Broeckhoven, C.; Martin, J.J. Presenile Alzheimer dementia characterized by amyloid angiopathy and large amyloid core type senile plaques in the APP 692Ala-->Gly mutation. Acta Neuropathol. 1998, 96, 253–260. [Google Scholar] [CrossRef]
  35. Braak, H.; Braak, E. Neuropathological stageing of Alzheimer-related changes. Acta Neuropathol. 1991, 82, 239–259. [Google Scholar] [CrossRef]
  36. Barrow, C.J.; Yasuda, A.; Kenny, P.T.; Zagorski, M.G. Solution conformations and aggregational properties of synthetic amyloid beta-peptides of Alzheimer’s disease. Analysis of circular dichroism spectra. J. Mol. Biol. 1992, 225, 1075–1093. [Google Scholar] [CrossRef]
  37. Fatafta, H.; Khaled, M.; Owen, M.C.; Sayyed-Ahmad, A.; Strodel, B. Amyloid-beta peptide dimers undergo a random coil to beta-sheet transition in the aqueous phase but not at the neuronal membrane. Proc. Natl. Acad. Sci. USA 2021, 118, e2106210118. [Google Scholar] [CrossRef]
  38. Ballatore, C.; Lee, V.M.; Trojanowski, J.Q. Tau-mediated neurodegeneration in Alzheimer’s disease and related disorders. Nat. Rev. Neurosci. 2007, 8, 663–672. [Google Scholar] [CrossRef]
  39. Goedert, M.; Spillantini, M.G.; Jakes, R.; Rutherford, D.; Crowther, R.A. Multiple isoforms of human microtubule-associated protein tau: Sequences and localization in neurofibrillary tangles of Alzheimer’s disease. Neuron 1989, 3, 519–526. [Google Scholar] [CrossRef]
  40. Polymeropoulos, M.H.; Lavedan, C.; Leroy, E.; Ide, S.E.; Dehejia, A.; Dutra, A.; Pike, B.; Root, H.; Rubenstein, J.; Boyer, R.; et al. Mutation in the alpha-synuclein gene identified in families with Parkinson’s disease. Science 1997, 276, 2045–2047. [Google Scholar] [CrossRef] [Green Version]
  41. Arai, T.; Hasegawa, M.; Akiyama, H.; Ikeda, K.; Nonaka, T.; Mori, H.; Mann, D.; Tsuchiya, K.; Yoshida, M.; Hashizume, Y.; et al. TDP-43 is a component of ubiquitin-positive tau-negative inclusions in frontotemporal lobar degeneration and amyotrophic lateral sclerosis. Biochem. Biophys. Res. Commun. 2006, 351, 602–611. [Google Scholar] [CrossRef]
  42. Neumann, M.; Sampathu, D.M.; Kwong, L.K.; Truax, A.C.; Micsenyi, M.C.; Chou, T.T.; Bruce, J.; Schuck, T.; Grossman, M.; Clark, C.M.; et al. Ubiquitinated TDP-43 in frontotemporal lobar degeneration and amyotrophic lateral sclerosis. Science 2006, 314, 130–133. [Google Scholar] [CrossRef] [PubMed]
  43. Ross, C.A.; Poirier, M.A. Protein aggregation and neurodegenerative disease. Nat. Med. 2004, 10 (Suppl. S7), S10–S17. [Google Scholar] [CrossRef] [PubMed]
  44. Soto, C. Unfolding the role of protein misfolding in neurodegenerative diseases. Nat. Rev. Neurosci. 2003, 4, 49–60. [Google Scholar] [CrossRef]
  45. Wu, L.; Rosa-Neto, P.; Hsiung, G.Y.; Sadovnick, A.D.; Masellis, M.; Black, S.E.; Jia, J.; Gauthier, S. Early-onset familial Alzheimer’s disease (EOFAD). Can. J. Neurol. Sci. 2012, 39, 436–445. [Google Scholar] [CrossRef] [PubMed]
  46. Tran, J.; Anastacio, H.; Bardy, C. Genetic predispositions of Parkinson’s disease revealed in patient-derived brain cells. NPJ Park. Dis. 2020, 6, 8. [Google Scholar] [CrossRef]
  47. Emery, A.E.; Holloway, S. Familial motor neuron diseases. Adv. Neurol. 1982, 36, 139–147. [Google Scholar]
  48. Sergeant, N.; David, J.P.; Lefranc, D.; Vermersch, P.; Wattez, A.; Delacourte, A. Different distribution of phosphorylated tau protein isoforms in Alzheimer’s and Pick’s diseases. FEBS Lett. 1997, 412, 578–582. [Google Scholar] [CrossRef]
  49. Stanford, P.M.; Halliday, G.M.; Brooks, W.S.; Kwok, J.B.; Storey, C.E.; Creasey, H.; Morris, J.G.; Fulham, M.J.; Schofield, P.R. Progressive supranuclear palsy pathology caused by a novel silent mutation in exon 10 of the tau gene: Expansion of the disease phenotype caused by tau gene mutations. Brain 2000, 123 Pt 5, 880–893. [Google Scholar] [CrossRef]
  50. Buratti, E. Functional Significance of TDP-43 Mutations in Disease. Adv. Genet. 2015, 91, 1–53. [Google Scholar] [CrossRef]
  51. Gamblin, T.C.; Chen, F.; Zambrano, A.; Abraha, A.; Lagalwar, S.; Guillozet, A.L.; Lu, M.; Fu, Y.; Garcia-Sierra, F.; LaPointe, N.; et al. Caspase cleavage of tau: Linking amyloid and neurofibrillary tangles in Alzheimer’s disease. Proc. Natl. Acad. Sci. USA 2003, 100, 10032–10037. [Google Scholar] [CrossRef] [Green Version]
  52. Bertram, L.; Tanzi, R.E. The genetics of Alzheimer’s disease. Prog. Mol. Biol. Transl. Sci. 2012, 107, 79–100. [Google Scholar] [CrossRef]
  53. Sorrentino, Z.A.; Giasson, B.I. The emerging role of alpha-synuclein truncation in aggregation and disease. J. Biol. Chem. 2020, 295, 10224–10244. [Google Scholar] [CrossRef]
  54. Cruts, M.; Theuns, J.; Van Broeckhoven, C. Locus-specific mutation databases for neurodegenerative brain diseases. Hum. Mutat. 2012, 33, 1340–1344. [Google Scholar] [CrossRef]
  55. Nilsberth, C.; Westlind-Danielsson, A.; Eckman, C.B.; Condron, M.M.; Axelman, K.; Forsell, C.; Stenh, C.; Luthman, J.; Teplow, D.B.; Younkin, S.G.; et al. The ‘Arctic’ APP mutation (E693G) causes Alzheimer’s disease by enhanced Abeta protofibril formation. Nat. Neurosci. 2001, 4, 887–893. [Google Scholar] [CrossRef]
  56. Schoch, K.M.; DeVos, S.L.; Miller, R.L.; Chun, S.J.; Norrbom, M.; Wozniak, D.F.; Dawson, H.N.; Bennett, C.F.; Rigo, F.; Miller, T.M. Increased 4R-Tau Induces Pathological Changes in a Human-Tau Mouse Model. Neuron 2016, 90, 941–947. [Google Scholar] [CrossRef]
  57. Johnson, B.S.; Snead, D.; Lee, J.J.; McCaffery, J.M.; Shorter, J.; Gitler, A.D. TDP-43 is intrinsically aggregation-prone, and amyotrophic lateral sclerosis-linked mutations accelerate aggregation and increase toxicity. J. Biol. Chem. 2009, 284, 20329–20339. [Google Scholar] [CrossRef]
  58. Selkoe, D.J. Alzheimer’s disease: Genes, proteins, and therapy. Physiol. Rev. 2001, 81, 741–766. [Google Scholar] [CrossRef]
  59. Hooli, B.V.; Mohapatra, G.; Mattheisen, M.; Parrado, A.R.; Roehr, J.T.; Shen, Y.; Gusella, J.F.; Moir, R.; Saunders, A.J.; Lange, C.; et al. Role of common and rare APP DNA sequence variants in Alzheimer disease. Neurology 2012, 78, 1250–1257. [Google Scholar] [CrossRef]
  60. Sahlin, C.; Lord, A.; Magnusson, K.; Englund, H.; Almeida, C.G.; Greengard, P.; Nyberg, F.; Gouras, G.K.; Lannfelt, L.; Nilsson, L.N. The Arctic Alzheimer mutation favors intracellular amyloid-beta production by making amyloid precursor protein less available to alpha-secretase. J. Neurochem. 2007, 101, 854–862. [Google Scholar] [CrossRef]
  61. Johansson, A.S.; Berglind-Dehlin, F.; Karlsson, G.; Edwards, K.; Gellerfors, P.; Lannfelt, L. Physiochemical characterization of the Alzheimer’s disease-related peptides A beta 1-42Arctic and A beta 1-42wt. FEBS J. 2006, 273, 2618–2630. [Google Scholar] [CrossRef] [PubMed]
  62. Novak, M.; Jakes, R.; Edwards, P.C.; Milstein, C.; Wischik, C.M. Difference between the tau protein of Alzheimer paired helical filament core and normal tau revealed by epitope analysis of monoclonal antibodies 423 and 7.51. Proc. Natl. Acad. Sci. USA 1991, 88, 5837–5841. [Google Scholar] [CrossRef] [PubMed]
  63. Alonso, A.D.; Cohen, L.S.; Corbo, C.; Morozova, V.; ElIdrissi, A.; Phillips, G.; Kleiman, F.E. Hyperphosphorylation of Tau Associates With Changes in Its Function Beyond Microtubule Stability. Front. Cell. Neurosci. 2018, 12, 338. [Google Scholar] [CrossRef] [PubMed]
  64. Poorkaj, P.; Bird, T.D.; Wijsman, E.; Nemens, E.; Garruto, R.M.; Anderson, L.; Andreadis, A.; Wiederholt, W.C.; Raskind, M.; Schellenberg, G.D. Tau is a candidate gene for chromosome 17 frontotemporal dementia. Ann. Neurol. 1998, 43, 815–825. [Google Scholar] [CrossRef] [PubMed]
  65. Hutton, M.; Lendon, C.L.; Rizzu, P.; Baker, M.; Froelich, S.; Houlden, H.; Pickering-Brown, S.; Chakraverty, S.; Isaacs, A.; Grover, A.; et al. Association of missense and 5′-splice-site mutations in tau with the inherited dementia FTDP-17. Nature 1998, 393, 702–705. [Google Scholar] [CrossRef]
  66. Spillantini, M.G.; Murrell, J.R.; Goedert, M.; Farlow, M.; Klug, A.; Ghetti, B. Mutations in the tau gene (MAPT) in FTDP-17: The family with Multiple System Tauopathy with Presenile Dementia (MSTD). J. Alzheimers Dis. 2006, 9, 373–380. [Google Scholar] [CrossRef]
  67. Vinueza-Gavilanes, R.; Inigo-Marco, I.; Larrea, L.; Lasa, M.; Carte, B.; Santamaria, E.; Fernandez-Irigoyen, J.; Bugallo, R.; Aragon, T.; Aldabe, R.; et al. N-terminal acetylation mutants affect alpha-synuclein stability, protein levels and neuronal toxicity. Neurobiol. Dis. 2020, 137, 104781. [Google Scholar] [CrossRef]
  68. Zarranz, J.J.; Alegre, J.; Gomez-Esteban, J.C.; Lezcano, E.; Ros, R.; Ampuero, I.; Vidal, L.; Hoenicka, J.; Rodriguez, O.; Atares, B.; et al. The new mutation, E46K, of alpha-synuclein causes Parkinson and Lewy body dementia. Ann. Neurol. 2004, 55, 164–173. [Google Scholar] [CrossRef]
  69. Kruger, R.; Kuhn, W.; Muller, T.; Woitalla, D.; Graeber, M.; Kosel, S.; Przuntek, H.; Epplen, J.T.; Schols, L.; Riess, O. Ala30Pro mutation in the gene encoding alpha-synuclein in Parkinson’s disease. Nat. Genet. 1998, 18, 106–108. [Google Scholar] [CrossRef]
  70. Robotta, M.; Cattani, J.; Martins, J.C.; Subramaniam, V.; Drescher, M. Alpha-Synuclein Disease Mutations Are Structurally Defective and Locally Affect Membrane Binding. J. Am. Chem. Soc. 2017, 139, 4254–4257. [Google Scholar] [CrossRef]
  71. Ferreon, A.C.; Moran, C.R.; Ferreon, J.C.; Deniz, A.A. Alteration of the alpha-synuclein folding landscape by a mutation related to Parkinson’s disease. Angew. Chem. Int. Ed. Engl. 2010, 49, 3469–3472. [Google Scholar] [CrossRef] [Green Version]
  72. Li, J.; Uversky, V.N.; Fink, A.L. Effect of familial Parkinson’s disease point mutations A30P and A53T on the structural properties, aggregation, and fibrillation of human alpha-synuclein. Biochemistry 2001, 40, 11604–11613. [Google Scholar] [CrossRef]
  73. Lv, Z.; Krasnoslobodtsev, A.V.; Zhang, Y.; Ysselstein, D.; Rochet, J.C.; Blanchard, S.C.; Lyubchenko, Y.L. Direct Detection of alpha-Synuclein Dimerization Dynamics: Single-Molecule Fluorescence Analysis. Biophys. J. 2015, 108, 2038–2047. [Google Scholar] [CrossRef]
  74. Kumar, S.; Jangir, D.K.; Kumar, R.; Kumari, M.; Bhavesh, N.S.; Maiti, T.K. Role of Sporadic Parkinson Disease Associated Mutations A18T and A29S in Enhanced alpha-Synuclein Fibrillation and Cytotoxicity. ACS Chem. Neurosci. 2018, 9, 230–240. [Google Scholar] [CrossRef]
  75. Sreedharan, J.; Blair, I.P.; Tripathi, V.B.; Hu, X.; Vance, C.; Rogelj, B.; Ackerley, S.; Durnall, J.C.; Williams, K.L.; Buratti, E.; et al. TDP-43 mutations in familial and sporadic amyotrophic lateral sclerosis. Science 2008, 319, 1668–1672. [Google Scholar] [CrossRef]
  76. Chen, A.K.; Lin, R.Y.; Hsieh, E.Z.; Tu, P.H.; Chen, R.P.; Liao, T.Y.; Chen, W.; Wang, C.H.; Huang, J.J. Induction of amyloid fibrils by the C-terminal fragments of TDP-43 in amyotrophic lateral sclerosis. J. Am. Chem. Soc. 2010, 132, 1186–1187. [Google Scholar] [CrossRef]
  77. Guo, W.; Chen, Y.; Zhou, X.; Kar, A.; Ray, P.; Chen, X.; Rao, E.J.; Yang, M.; Ye, H.; Zhu, L.; et al. An ALS-associated mutation affecting TDP-43 enhances protein aggregation, fibril formation and neurotoxicity. Nat. Struct. Mol. Biol. 2011, 18, 822–830. [Google Scholar] [CrossRef]
  78. Weggen, S.; Beher, D. Molecular consequences of amyloid precursor protein and presenilin mutations causing autosomal-dominant Alzheimer’s disease. Alzheimers Res. Ther. 2012, 4, 9. [Google Scholar] [CrossRef]
  79. Alonso Adel, C.; Mederlyova, A.; Novak, M.; Grundke-Iqbal, I.; Iqbal, K. Promotion of hyperphosphorylation by frontotemporal dementia tau mutations. J. Biol. Chem. 2004, 279, 34873–34881. [Google Scholar] [CrossRef]
  80. Goedert, M. Tau protein and neurodegeneration. Semin. Cell Dev. Biol. 2004, 15, 45–49. [Google Scholar] [CrossRef]
  81. Pasanen, P.; Myllykangas, L.; Siitonen, M.; Raunio, A.; Kaakkola, S.; Lyytinen, J.; Tienari, P.J.; Poyhonen, M.; Paetau, A. Novel alpha-synuclein mutation A53E associated with atypical multiple system atrophy and Parkinson’s disease-type pathology. Neurobiol. Aging 2014, 35, 2180.e1–2180.e5. [Google Scholar] [CrossRef] [PubMed]
  82. Guan, Y.; Zhao, X.; Liu, F.; Yan, S.; Wang, Y.; Du, C.; Cui, X.; Li, R.; Zhang, C.X. Pathogenic Mutations Differentially Regulate Cell-to-Cell Transmission of alpha-Synuclein. Front. Cell. Neurosci. 2020, 14, 159. [Google Scholar] [CrossRef] [PubMed]
  83. Kabashi, E.; Valdmanis, P.N.; Dion, P.; Spiegelman, D.; McConkey, B.J.; Vande Velde, C.; Bouchard, J.P.; Lacomblez, L.; Pochigaeva, K.; Salachas, F.; et al. TARDBP mutations in individuals with sporadic and familial amyotrophic lateral sclerosis. Nat. Genet. 2008, 40, 572–574. [Google Scholar] [CrossRef] [PubMed]
  84. Nonaka, T.; Kametani, F.; Arai, T.; Akiyama, H.; Hasegawa, M. Truncation and pathogenic mutations facilitate the formation of intracellular aggregates of TDP-43. Hum. Mol. Genet. 2009, 18, 3353–3364. [Google Scholar] [CrossRef] [PubMed]
  85. Bah, A.; Forman-Kay, J.D. Modulation of Intrinsically Disordered Protein Function by Post-translational Modifications. J. Biol. Chem. 2016, 291, 6696–6705. [Google Scholar] [CrossRef]
  86. Darling, A.L.; Uversky, V.N. Intrinsic Disorder and Posttranslational Modifications: The Darker Side of the Biological Dark Matter. Front. Genet. 2018, 9, 158. [Google Scholar] [CrossRef]
  87. Bramblett, G.T.; Goedert, M.; Jakes, R.; Merrick, S.E.; Trojanowski, J.Q.; Lee, V.M. Abnormal tau phosphorylation at Ser396 in Alzheimer’s disease recapitulates development and contributes to reduced microtubule binding. Neuron 1993, 10, 1089–1099. [Google Scholar] [CrossRef]
  88. Kumar, S.; Walter, J. Phosphorylation of amyloid beta (Abeta) peptides—A trigger for formation of toxic aggregates in Alzheimer’s disease. Aging 2011, 3, 803–812. [Google Scholar] [CrossRef]
  89. Gerson, J.E.; Sengupta, U.; Lasagna-Reeves, C.A.; Guerrero-Munoz, M.J.; Troncoso, J.; Kayed, R. Characterization of tau oligomeric seeds in progressive supranuclear palsy. Acta Neuropathol. Commun. 2014, 2, 73. [Google Scholar] [CrossRef]
  90. Kumar, S.; Rezaei-Ghaleh, N.; Terwel, D.; Thal, D.R.; Richard, M.; Hoch, M.; Mc Donald, J.M.; Wullner, U.; Glebov, K.; Heneka, M.T.; et al. Extracellular phosphorylation of the amyloid beta-peptide promotes formation of toxic aggregates during the pathogenesis of Alzheimer’s disease. EMBO J. 2011, 30, 2255–2265. [Google Scholar] [CrossRef]
  91. Rijal Upadhaya, A.; Kosterin, I.; Kumar, S.; von Arnim, C.A.; Yamaguchi, H.; Fandrich, M.; Walter, J.; Thal, D.R. Biochemical stages of amyloid-beta peptide aggregation and accumulation in the human brain and their association with symptomatic and pathologically preclinical Alzheimer’s disease. Brain 2014, 137, 887–903. [Google Scholar] [CrossRef]
  92. Iannuzzi, C.; Irace, G.; Sirangelo, I. Differential effects of glycation on protein aggregation and amyloid formation. Front. Mol. Biosci. 2014, 1, 9. [Google Scholar] [CrossRef]
  93. Vitek, M.P.; Bhattacharya, K.; Glendening, J.M.; Stopa, E.; Vlassara, H.; Bucala, R.; Manogue, K.; Cerami, A. Advanced glycation end products contribute to amyloidosis in Alzheimer disease. Proc. Natl. Acad. Sci. USA 1994, 91, 4766–4770. [Google Scholar] [CrossRef]
  94. Li, X.H.; Du, L.L.; Cheng, X.S.; Jiang, X.; Zhang, Y.; Lv, B.L.; Liu, R.; Wang, J.Z.; Zhou, X.W. Glycation exacerbates the neuronal toxicity of beta-amyloid. Cell Death Dis. 2013, 4, e673. [Google Scholar] [CrossRef]
  95. Halim, A.; Brinkmalm, G.; Ruetschi, U.; Westman-Brinkmalm, A.; Portelius, E.; Zetterberg, H.; Blennow, K.; Larson, G.; Nilsson, J. Site-specific characterization of threonine, serine, and tyrosine glycosylations of amyloid precursor protein/amyloid beta-peptides in human cerebrospinal fluid. Proc. Natl. Acad. Sci. USA 2011, 108, 11848–11853. [Google Scholar] [CrossRef]
  96. Liu, D.; Wei, Q.; Xia, W.; He, C.; Zhang, Q.; Huang, L.; Wang, X.; Sun, Y.; Ma, Y.; Zhang, X.; et al. O-Glycosylation Induces Amyloid-beta To Form New Fibril Polymorphs Vulnerable for Degradation. J. Am. Chem. Soc. 2021, 143, 20216–20223. [Google Scholar] [CrossRef]
  97. Cho, J.H.; Johnson, G.V. Glycogen synthase kinase 3beta phosphorylates tau at both primed and unprimed sites. Differential impact on microtubule binding. J. Biol. Chem. 2003, 278, 187–193. [Google Scholar] [CrossRef]
  98. Haase, C.; Stieler, J.T.; Arendt, T.; Holzer, M. Pseudophosphorylation of tau protein alters its ability for self-aggregation. J. Neurochem. 2004, 88, 1509–1520. [Google Scholar] [CrossRef]
  99. Min, S.W.; Cho, S.H.; Zhou, Y.; Schroeder, S.; Haroutunian, V.; Seeley, W.W.; Huang, E.J.; Shen, Y.; Masliah, E.; Mukherjee, C.; et al. Acetylation of tau inhibits its degradation and contributes to tauopathy. Neuron 2010, 67, 953–966. [Google Scholar] [CrossRef]
  100. Morris, M.; Knudsen, G.M.; Maeda, S.; Trinidad, J.C.; Ioanoviciu, A.; Burlingame, A.L.; Mucke, L. Tau post-translational modifications in wild-type and human amyloid precursor protein transgenic mice. Nat. Neurosci. 2015, 18, 1183–1189. [Google Scholar] [CrossRef]
  101. Cohen, T.J.; Guo, J.L.; Hurtado, D.E.; Kwong, L.K.; Mills, I.P.; Trojanowski, J.Q.; Lee, V.M. The acetylation of tau inhibits its function and promotes pathological tau aggregation. Nat. Commun. 2011, 2, 252. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  102. Irwin, D.J.; Cohen, T.J.; Grossman, M.; Arnold, S.E.; Xie, S.X.; Lee, V.M.; Trojanowski, J.Q. Acetylated tau, a novel pathological signature in Alzheimer’s disease and other tauopathies. Brain 2012, 135, 807–818. [Google Scholar] [CrossRef] [PubMed]
  103. Bancher, C.; Grundke-Iqbal, I.; Iqbal, K.; Fried, V.A.; Smith, H.T.; Wisniewski, H.M. Abnormal phosphorylation of tau precedes ubiquitination in neurofibrillary pathology of Alzheimer disease. Brain Res. 1991, 539, 11–18. [Google Scholar] [CrossRef] [PubMed]
  104. Morishima-Kawashima, M.; Hasegawa, M.; Takio, K.; Suzuki, M.; Titani, K.; Ihara, Y. Ubiquitin is conjugated with amino-terminally processed tau in paired helical filaments. Neuron 1993, 10, 1151–1160. [Google Scholar] [CrossRef]
  105. Tan, J.M.; Wong, E.S.; Kirkpatrick, D.S.; Pletnikova, O.; Ko, H.S.; Tay, S.P.; Ho, M.W.; Troncoso, J.; Gygi, S.P.; Lee, M.K.; et al. Lysine 63-linked ubiquitination promotes the formation and autophagic clearance of protein inclusions associated with neurodegenerative diseases. Hum. Mol. Genet. 2008, 17, 431–439. [Google Scholar] [CrossRef]
  106. Dickey, C.A.; Yue, M.; Lin, W.L.; Dickson, D.W.; Dunmore, J.H.; Lee, W.C.; Zehr, C.; West, G.; Cao, S.; Clark, A.M.; et al. Deletion of the ubiquitin ligase CHIP leads to the accumulation, but not the aggregation, of both endogenous phospho- and caspase-3-cleaved tau species. J. Neurosci. 2006, 26, 6985–6996. [Google Scholar] [CrossRef]
  107. Petrucelli, L.; Dickson, D.; Kehoe, K.; Taylor, J.; Snyder, H.; Grover, A.; De Lucia, M.; McGowan, E.; Lewis, J.; Prihar, G.; et al. CHIP and Hsp70 regulate tau ubiquitination, degradation and aggregation. Hum. Mol. Genet. 2004, 13, 703–714. [Google Scholar] [CrossRef]
  108. Fujiwara, H.; Hasegawa, M.; Dohmae, N.; Kawashima, A.; Masliah, E.; Goldberg, M.S.; Shen, J.; Takio, K.; Iwatsubo, T. alpha-Synuclein is phosphorylated in synucleinopathy lesions. Nat. Cell Biol. 2002, 4, 160–164. [Google Scholar] [CrossRef]
  109. Chen, L.; Periquet, M.; Wang, X.; Negro, A.; McLean, P.J.; Hyman, B.T.; Feany, M.B. Tyrosine and serine phosphorylation of alpha-synuclein have opposing effects on neurotoxicity and soluble oligomer formation. J. Clin. Investig. 2009, 119, 3257–3265. [Google Scholar] [CrossRef]
  110. Hasegawa, M.; Fujiwara, H.; Nonaka, T.; Wakabayashi, K.; Takahashi, H.; Lee, V.M.; Trojanowski, J.Q.; Mann, D.; Iwatsubo, T. Phosphorylated alpha-synuclein is ubiquitinated in alpha-synucleinopathy lesions. J. Biol. Chem. 2002, 277, 49071–49076. [Google Scholar] [CrossRef]
  111. Kim, Y.M.; Jang, W.H.; Quezado, M.M.; Oh, Y.; Chung, K.C.; Junn, E.; Mouradian, M.M. Proteasome inhibition induces alpha-synuclein SUMOylation and aggregate formation. J. Neurol. Sci. 2011, 307, 157–161. [Google Scholar] [CrossRef] [Green Version]
  112. Rott, R.; Szargel, R.; Shani, V.; Hamza, H.; Savyon, M.; Abd Elghani, F.; Bandopadhyay, R.; Engelender, S. SUMOylation and ubiquitination reciprocally regulate alpha-synuclein degradation and pathological aggregation. Proc. Natl. Acad. Sci. USA 2017, 114, 13176–13181. [Google Scholar] [CrossRef]
  113. Martinez-Vicente, M.; Talloczy, Z.; Kaushik, S.; Massey, A.C.; Mazzulli, J.; Mosharov, E.V.; Hodara, R.; Fredenburg, R.; Wu, D.C.; Follenzi, A.; et al. Dopamine-modified alpha-synuclein blocks chaperone-mediated autophagy. J. Clin. Investig. 2008, 118, 777–788. [Google Scholar] [CrossRef]
  114. Wang, Z.; Park, K.; Comer, F.; Hsieh-Wilson, L.C.; Saudek, C.D.; Hart, G.W. Site-specific GlcNAcylation of human erythrocyte proteins: Potential biomarker(s) for diabetes. Diabetes 2009, 58, 309–317. [Google Scholar] [CrossRef]
  115. Neumann, M.; Kwong, L.K.; Lee, E.B.; Kremmer, E.; Flatley, A.; Xu, Y.; Forman, M.S.; Troost, D.; Kretzschmar, H.A.; Trojanowski, J.Q.; et al. Phosphorylation of S409/410 of TDP-43 is a consistent feature in all sporadic and familial forms of TDP-43 proteinopathies. Acta Neuropathol. 2009, 117, 137–149. [Google Scholar] [CrossRef]
  116. Nonaka, T.; Arai, T.; Buratti, E.; Baralle, F.E.; Akiyama, H.; Hasegawa, M. Phosphorylated and ubiquitinated TDP-43 pathological inclusions in ALS and FTLD-U are recapitulated in SH-SY5Y cells. FEBS Lett. 2009, 583, 394–400. [Google Scholar] [CrossRef]
  117. Nonaka, T.; Suzuki, G.; Tanaka, Y.; Kametani, F.; Hirai, S.; Okado, H.; Miyashita, T.; Saitoe, M.; Akiyama, H.; Masai, H.; et al. Phosphorylation of TAR DNA-binding Protein of 43 kDa (TDP-43) by Truncated Casein Kinase 1delta Triggers Mislocalization and Accumulation of TDP-43. J. Biol. Chem. 2016, 291, 5473–5483. [Google Scholar] [CrossRef]
  118. Hebron, M.L.; Lonskaya, I.; Sharpe, K.; Weerasinghe, P.P.; Algarzae, N.K.; Shekoyan, A.R.; Moussa, C.E. Parkin ubiquitinates Tar-DNA binding protein-43 (TDP-43) and promotes its cytosolic accumulation via interaction with histone deacetylase 6 (HDAC6). J. Biol. Chem. 2013, 288, 4103–4115. [Google Scholar] [CrossRef]
  119. Seyfried, N.T.; Gozal, Y.M.; Dammer, E.B.; Xia, Q.; Duong, D.M.; Cheng, D.; Lah, J.J.; Levey, A.I.; Peng, J. Multiplex SILAC analysis of a cellular TDP-43 proteinopathy model reveals protein inclusions associated with SUMOylation and diverse polyubiquitin chains. Mol. Cell. Proteom. 2010, 9, 705–718. [Google Scholar] [CrossRef]
  120. Cohen, T.J.; Hwang, A.W.; Restrepo, C.R.; Yuan, C.X.; Trojanowski, J.Q.; Lee, V.M. An acetylation switch controls TDP-43 function and aggregation propensity. Nat. Commun. 2015, 6, 5845. [Google Scholar] [CrossRef]
  121. Kummer, M.P.; Heneka, M.T. Truncated and modified amyloid-beta species. Alzheimers Res. Ther. 2014, 6, 28. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  122. Manzanza, N.O.; Sedlackova, L.; Kalaria, R.N. Alpha-Synuclein Post-translational Modifications: Implications for Pathogenesis of Lewy Body Disorders. Front. Aging Neurosci. 2021, 13, 690293. [Google Scholar] [CrossRef] [PubMed]
  123. Alquezar, C.; Arya, S.; Kao, A.W. Tau Post-translational Modifications: Dynamic Transformers of Tau Function, Degradation, and Aggregation. Front. Neurol. 2020, 11, 595532. [Google Scholar] [CrossRef] [PubMed]
  124. Zhang, J.; Li, X.; Li, J.D. The Roles of Post-translational Modifications on alpha-Synuclein in the Pathogenesis of Parkinson’s Diseases. Front. Neurosci. 2019, 13, 381. [Google Scholar] [CrossRef] [PubMed]
  125. Buratti, E. TDP-43 post-translational modifications in health and disease. Expert Opin. Ther. Targets 2018, 22, 279–293. [Google Scholar] [CrossRef]
  126. Kovacech, B.; Skrabana, R.; Novak, M. Transition of tau protein from disordered to misordered in Alzheimer’s disease. Neurodegener. Dis. 2010, 7, 24–27. [Google Scholar] [CrossRef]
  127. Younkin, S.G. The role of A beta 42 in Alzheimer’s disease. J. Physiol. Paris 1998, 92, 289–292. [Google Scholar] [CrossRef]
  128. Ma, L.; Yang, C.; Zhang, X.; Li, Y.; Wang, S.; Zheng, L.; Huang, K. C-terminal truncation exacerbates the aggregation and cytotoxicity of alpha-Synuclein: A vicious cycle in Parkinson’s disease. Biochim. Biophys. Acta Mol. Basis Dis. 2018, 1864, 3714–3725. [Google Scholar] [CrossRef]
  129. O’Brien, R.J.; Wong, P.C. Amyloid precursor protein processing and Alzheimer’s disease. Annu. Rev. Neurosci. 2011, 34, 185–204. [Google Scholar] [CrossRef]
  130. Zhang, Y.W.; Thompson, R.; Zhang, H.; Xu, H. APP processing in Alzheimer’s disease. Mol. Brain 2011, 4, 3. [Google Scholar] [CrossRef]
  131. Bibl, M.; Mollenhauer, B.; Esselmann, H.; Lewczuk, P.; Klafki, H.W.; Sparbier, K.; Smirnov, A.; Cepek, L.; Trenkwalder, C.; Ruther, E.; et al. CSF amyloid-beta-peptides in Alzheimer’s disease, dementia with Lewy bodies and Parkinson’s disease dementia. Brain 2006, 129, 1177–1187. [Google Scholar] [CrossRef]
  132. Bibl, M.; Mollenhauer, B.; Lewczuk, P.; Esselmann, H.; Wolf, S.; Trenkwalder, C.; Otto, M.; Stiens, G.; Ruther, E.; Kornhuber, J.; et al. Validation of amyloid-beta peptides in CSF diagnosis of neurodegenerative dementias. Mol. Psychiatry 2007, 12, 671–680. [Google Scholar] [CrossRef]
  133. Welge, V.; Fiege, O.; Lewczuk, P.; Mollenhauer, B.; Esselmann, H.; Klafki, H.W.; Wolf, S.; Trenkwalder, C.; Otto, M.; Kornhuber, J.; et al. Combined CSF tau, p-tau181 and amyloid-beta 38/40/42 for diagnosing Alzheimer’s disease. J. Neural. Transm. 2009, 116, 203–212. [Google Scholar] [CrossRef]
  134. Tang, W.; Huang, Q.; Wang, Y.; Wang, Z.Y.; Yao, Y.Y. Assessment of CSF Abeta42 as an aid to discriminating Alzheimer’s disease from other dementias and mild cognitive impairment: A meta-analysis of 50 studies. J. Neurol. Sci. 2014, 345, 26–36. [Google Scholar] [CrossRef]
  135. Nirmalraj, P.N.; List, J.; Battacharya, S.; Howe, G.; Xu, L.; Thompson, D.; Mayer, M. Complete aggregation pathway of amyloid beta (1-40) and (1-42) resolved on an atomically clean interface. Sci. Adv. 2020, 6, eaaz6014. [Google Scholar] [CrossRef]
  136. Asami-Odaka, A.; Ishibashi, Y.; Kikuchi, T.; Kitada, C.; Suzuki, N. Long amyloid beta-protein secreted from wild-type human neuroblastoma IMR-32 cells. Biochemistry 1995, 34, 10272–10278. [Google Scholar] [CrossRef]
  137. Zhao, X.; Kotilinek, L.A.; Smith, B.; Hlynialuk, C.; Zahs, K.; Ramsden, M.; Cleary, J.; Ashe, K.H. Caspase-2 cleavage of tau reversibly impairs memory. Nat. Med. 2016, 22, 1268–1276. [Google Scholar] [CrossRef]
  138. Rissman, R.A.; Poon, W.W.; Blurton-Jones, M.; Oddo, S.; Torp, R.; Vitek, M.P.; LaFerla, F.M.; Rohn, T.T.; Cotman, C.W. Caspase-cleavage of tau is an early event in Alzheimer disease tangle pathology. J. Clin. Investig. 2004, 114, 121–130. [Google Scholar] [CrossRef]
  139. Means, J.C.; Gerdes, B.C.; Kaja, S.; Sumien, N.; Payne, A.J.; Stark, D.A.; Borden, P.K.; Price, J.L.; Koulen, P. Caspase-3-Dependent Proteolytic Cleavage of Tau Causes Neurofibrillary Tangles and Results in Cognitive Impairment During Normal Aging. Neurochem. Res. 2016, 41, 2278–2288. [Google Scholar] [CrossRef]
  140. Guo, H.; Albrecht, S.; Bourdeau, M.; Petzke, T.; Bergeron, C.; LeBlanc, A.C. Active caspase-6 and caspase-6-cleaved tau in neuropil threads, neuritic plaques, and neurofibrillary tangles of Alzheimer’s disease. Am. J. Pathol. 2004, 165, 523–531. [Google Scholar] [CrossRef]
  141. Ramcharitar, J.; Albrecht, S.; Afonso, V.M.; Kaushal, V.; Bennett, D.A.; Leblanc, A.C. Cerebrospinal fluid tau cleaved by caspase-6 reflects brain levels and cognition in aging and Alzheimer disease. J. Neuropathol. Exp. Neurol. 2013, 72, 824–832. [Google Scholar] [CrossRef] [PubMed]
  142. Park, S.Y.; Ferreira, A. The generation of a 17 kDa neurotoxic fragment: An alternative mechanism by which tau mediates beta-amyloid-induced neurodegeneration. J. Neurosci. 2005, 25, 5365–5375. [Google Scholar] [CrossRef] [PubMed]
  143. Matsumoto, S.E.; Motoi, Y.; Ishiguro, K.; Tabira, T.; Kametani, F.; Hasegawa, M.; Hattori, N. The twenty-four KDa C-terminal tau fragment increases with aging in tauopathy mice: Implications of prion-like properties. Hum. Mol. Genet. 2015, 24, 6403–6416. [Google Scholar] [CrossRef] [PubMed]
  144. Mishizen-Eberz, A.J.; Guttmann, R.P.; Giasson, B.I.; Day, G.A., 3rd; Hodara, R.; Ischiropoulos, H.; Lee, V.M.; Trojanowski, J.Q.; Lynch, D.R. Distinct cleavage patterns of normal and pathologic forms of alpha-synuclein by calpain I in vitro. J. Neurochem. 2003, 86, 836–847. [Google Scholar] [CrossRef] [PubMed]
  145. Li, W.; West, N.; Colla, E.; Pletnikova, O.; Troncoso, J.C.; Marsh, L.; Dawson, T.M.; Jakala, P.; Hartmann, T.; Price, D.L.; et al. Aggregation promoting C-terminal truncation of alpha-synuclein is a normal cellular process and is enhanced by the familial Parkinson’s disease-linked mutations. Proc. Natl. Acad. Sci. USA 2005, 102, 2162–2167. [Google Scholar] [CrossRef]
  146. Liu, C.W.; Giasson, B.I.; Lewis, K.A.; Lee, V.M.; Demartino, G.N.; Thomas, P.J. A precipitating role for truncated alpha-synuclein and the proteasome in alpha-synuclein aggregation: Implications for pathogenesis of Parkinson disease. J. Biol. Chem. 2005, 280, 22670–22678. [Google Scholar] [CrossRef]
  147. Tofaris, G.K.; Garcia Reitbock, P.; Humby, T.; Lambourne, S.L.; O’Connell, M.; Ghetti, B.; Gossage, H.; Emson, P.C.; Wilkinson, L.S.; Goedert, M.; et al. Pathological changes in dopaminergic nerve cells of the substantia nigra and olfactory bulb in mice transgenic for truncated human alpha-synuclein(1-120): Implications for Lewy body disorders. J. Neurosci. 2006, 26, 3942–3950. [Google Scholar] [CrossRef]
  148. Wang, W.; Nguyen, L.T.; Burlak, C.; Chegini, F.; Guo, F.; Chataway, T.; Ju, S.; Fisher, O.S.; Miller, D.W.; Datta, D.; et al. Caspase-1 causes truncation and aggregation of the Parkinson’s disease-associated protein alpha-synuclein. Proc. Natl. Acad. Sci. USA 2016, 113, 9587–9592. [Google Scholar] [CrossRef]
  149. Zhang, Y.J.; Xu, Y.F.; Cook, C.; Gendron, T.F.; Roettges, P.; Link, C.D.; Lin, W.L.; Tong, J.; Castanedes-Casey, M.; Ash, P.; et al. Aberrant cleavage of TDP-43 enhances aggregation and cellular toxicity. Proc. Natl. Acad. Sci. USA 2009, 106, 7607–7612. [Google Scholar] [CrossRef]
  150. Zhang, Y.J.; Xu, Y.F.; Dickey, C.A.; Buratti, E.; Baralle, F.; Bailey, R.; Pickering-Brown, S.; Dickson, D.; Petrucelli, L. Progranulin mediates caspase-dependent cleavage of TAR DNA binding protein-43. J. Neurosci. 2007, 27, 10530–10534. [Google Scholar] [CrossRef]
  151. Jiang, L.L.; Che, M.X.; Zhao, J.; Zhou, C.J.; Xie, M.Y.; Li, H.Y.; He, J.H.; Hu, H.Y. Structural transformation of the amyloidogenic core region of TDP-43 protein initiates its aggregation and cytoplasmic inclusion. J. Biol. Chem. 2013, 288, 19614–19624. [Google Scholar] [CrossRef] [Green Version]
  152. Gu, J.; Xu, W.; Jin, N.; Li, L.; Zhou, Y.; Chu, D.; Gong, C.X.; Iqbal, K.; Liu, F. Truncation of Tau selectively facilitates its pathological activities. J. Biol. Chem. 2020, 295, 13812–13828. [Google Scholar] [CrossRef]
  153. Nishimoto, Y.; Ito, D.; Yagi, T.; Nihei, Y.; Tsunoda, Y.; Suzuki, N. Characterization of alternative isoforms and inclusion body of the TAR DNA-binding protein-43. J. Biol. Chem. 2010, 285, 608–619. [Google Scholar] [CrossRef]
  154. Berning, B.A.; Walker, A.K. The Pathobiology of TDP-43 C-Terminal Fragments in ALS and FTLD. Front. Neurosci. 2019, 13, 335. [Google Scholar] [CrossRef]
  155. Caughey, B.; Lansbury, P.T. Protofibrils, pores, fibrils, and neurodegeneration: Separating the responsible protein aggregates from the innocent bystanders. Annu. Rev. Neurosci. 2003, 26, 267–298. [Google Scholar] [CrossRef]
  156. Gadad, B.S.; Britton, G.B.; Rao, K.S. Targeting oligomers in neurodegenerative disorders: Lessons from alpha-synuclein, tau, and amyloid-beta peptide. J. Alzheimers Dis. 2011, 24 (Suppl. S2), 223–232. [Google Scholar] [CrossRef]
  157. Glabe, C.G. Common mechanisms of amyloid oligomer pathogenesis in degenerative disease. Neurobiol. Aging 2006, 27, 570–575. [Google Scholar] [CrossRef]
  158. Rambaran, R.N.; Serpell, L.C. Amyloid fibrils: Abnormal protein assembly. Prion 2008, 2, 112–117. [Google Scholar] [CrossRef]
  159. Benilova, I.; Karran, E.; De Strooper, B. The toxic Abeta oligomer and Alzheimer’s disease: An emperor in need of clothes. Nat. Neurosci. 2012, 15, 349–357. [Google Scholar] [CrossRef]
  160. Breydo, L.; Uversky, V.N. Structural, morphological, and functional diversity of amyloid oligomers. FEBS Lett. 2015, 589, 2640–2648. [Google Scholar] [CrossRef]
  161. Lesne, S.E. Toxic oligomer species of amyloid-beta in Alzheimer’s disease, a timing issue. Swiss Med. Wkly. 2014, 144, w14021. [Google Scholar] [CrossRef] [PubMed]
  162. Lambert, M.P.; Barlow, A.K.; Chromy, B.A.; Edwards, C.; Freed, R.; Liosatos, M.; Morgan, T.E.; Rozovsky, I.; Trommer, B.; Viola, K.L.; et al. Diffusible, nonfibrillar ligands derived from Abeta1-42 are potent central nervous system neurotoxins. Proc. Natl. Acad. Sci. USA 1998, 95, 6448–6453. [Google Scholar] [CrossRef] [PubMed]
  163. Matsumura, S.; Shinoda, K.; Yamada, M.; Yokojima, S.; Inoue, M.; Ohnishi, T.; Shimada, T.; Kikuchi, K.; Masui, D.; Hashimoto, S.; et al. Two distinct amyloid beta-protein (Abeta) assembly pathways leading to oligomers and fibrils identified by combined fluorescence correlation spectroscopy, morphology, and toxicity analyses. J. Biol. Chem. 2011, 286, 11555–11562. [Google Scholar] [CrossRef]
  164. Lacor, P.N.; Buniel, M.C.; Chang, L.; Fernandez, S.J.; Gong, Y.; Viola, K.L.; Lambert, M.P.; Velasco, P.T.; Bigio, E.H.; Finch, C.E.; et al. Synaptic targeting by Alzheimer’s-related amyloid beta oligomers. J. Neurosci. 2004, 24, 10191–10200. [Google Scholar] [CrossRef] [PubMed]
  165. Lacor, P.N.; Buniel, M.C.; Furlow, P.W.; Clemente, A.S.; Velasco, P.T.; Wood, M.; Viola, K.L.; Klein, W.L. Abeta oligomer-induced aberrations in synapse composition, shape, and density provide a molecular basis for loss of connectivity in Alzheimer’s disease. J. Neurosci. 2007, 27, 796–807. [Google Scholar] [CrossRef]
  166. Velasco, P.T.; Heffern, M.C.; Sebollela, A.; Popova, I.A.; Lacor, P.N.; Lee, K.B.; Sun, X.; Tiano, B.N.; Viola, K.L.; Eckermann, A.L.; et al. Synapse-binding subpopulations of Abeta oligomers sensitive to peptide assembly blockers and scFv antibodies. ACS Chem. Neurosci. 2012, 3, 972–981. [Google Scholar] [CrossRef]
  167. Liu, P.; Reed, M.N.; Kotilinek, L.A.; Grant, M.K.; Forster, C.L.; Qiang, W.; Shapiro, S.L.; Reichl, J.H.; Chiang, A.C.; Jankowsky, J.L.; et al. Quaternary Structure Defines a Large Class of Amyloid-beta Oligomers Neutralized by Sequestration. Cell Rep. 2015, 11, 1760–1771. [Google Scholar] [CrossRef]
  168. Hartley, D.M.; Walsh, D.M.; Ye, C.P.; Diehl, T.; Vasquez, S.; Vassilev, P.M.; Teplow, D.B.; Selkoe, D.J. Protofibrillar intermediates of amyloid beta-protein induce acute electrophysiological changes and progressive neurotoxicity in cortical neurons. J. Neurosci. 1999, 19, 8876–8884. [Google Scholar] [CrossRef]
  169. Lord, A.; Englund, H.; Soderberg, L.; Tucker, S.; Clausen, F.; Hillered, L.; Gordon, M.; Morgan, D.; Lannfelt, L.; Pettersson, F.E.; et al. Amyloid-beta protofibril levels correlate with spatial learning in Arctic Alzheimer’s disease transgenic mice. FEBS J. 2009, 276, 995–1006. [Google Scholar] [CrossRef]
  170. Lasagna-Reeves, C.A.; Kayed, R. Astrocytes contain amyloid-beta annular protofibrils in Alzheimer’s disease brains. FEBS Lett. 2011, 585, 3052–3057. [Google Scholar] [CrossRef]
  171. Gerson, J.; Castillo-Carranza, D.L.; Sengupta, U.; Bodani, R.; Prough, D.S.; DeWitt, D.S.; Hawkins, B.E.; Kayed, R. Tau Oligomers Derived from Traumatic Brain Injury Cause Cognitive Impairment and Accelerate Onset of Pathology in Htau Mice. J. Neurotrauma 2016, 33, 2034–2043. [Google Scholar] [CrossRef]
  172. Hawkins, B.E.; Krishnamurthy, S.; Castillo-Carranza, D.L.; Sengupta, U.; Prough, D.S.; Jackson, G.R.; DeWitt, D.S.; Kayed, R. Rapid accumulation of endogenous tau oligomers in a rat model of traumatic brain injury: Possible link between traumatic brain injury and sporadic tauopathies. J. Biol. Chem. 2013, 288, 17042–17050. [Google Scholar] [CrossRef]
  173. Himmelstein, D.S.; Ward, S.M.; Lancia, J.K.; Patterson, K.R.; Binder, L.I. Tau as a therapeutic target in neurodegenerative disease. Pharmacol. Ther. 2012, 136, 8–22. [Google Scholar] [CrossRef]
  174. Takashima, A. Tauopathies and tau oligomers. J. Alzheimers Dis. 2013, 37, 565–568. [Google Scholar] [CrossRef]
  175. Maeda, S.; Sahara, N.; Saito, Y.; Murayama, S.; Ikai, A.; Takashima, A. Increased levels of granular tau oligomers: An early sign of brain aging and Alzheimer’s disease. Neurosci. Res. 2006, 54, 197–201. [Google Scholar] [CrossRef]
  176. Goldberg, M.S.; Lansbury, P.T., Jr. Is there a cause-and-effect relationship between alpha-synuclein fibrillization and Parkinson’s disease? Nat. Cell Biol. 2000, 2, E115–E119. [Google Scholar] [CrossRef]
  177. Conway, K.A.; Lee, S.J.; Rochet, J.C.; Ding, T.T.; Williamson, R.E.; Lansbury, P.T., Jr. Acceleration of oligomerization, not fibrillization, is a shared property of both alpha-synuclein mutations linked to early-onset Parkinson’s disease: Implications for pathogenesis and therapy. Proc. Natl. Acad. Sci. USA 2000, 97, 571–576. [Google Scholar] [CrossRef]
  178. Moosa, M.M.; Ferreon, J.C.; Ferreon, A.C.M. Single-Molecule FRET Detection of Early-Stage Conformations in alpha-Synuclein Aggregation. Methods Mol. Biol. 2019, 1948, 221–233. [Google Scholar] [CrossRef]
  179. Dearborn, A.D.; Wall, J.S.; Cheng, N.; Heymann, J.B.; Kajava, A.V.; Varkey, J.; Langen, R.; Steven, A.C. alpha-Synuclein Amyloid Fibrils with Two Entwined, Asymmetrically Associated Protofibrils. J. Biol. Chem. 2016, 291, 2310–2318. [Google Scholar] [CrossRef]
  180. Lashuel, H.A.; Petre, B.M.; Wall, J.; Simon, M.; Nowak, R.J.; Walz, T.; Lansbury, P.T., Jr. Alpha-synuclein, especially the Parkinson’s disease-associated mutants, forms pore-like annular and tubular protofibrils. J. Mol. Biol. 2002, 322, 1089–1102. [Google Scholar] [CrossRef]
  181. Zhang, Y.J.; Caulfield, T.; Xu, Y.F.; Gendron, T.F.; Hubbard, J.; Stetler, C.; Sasaguri, H.; Whitelaw, E.C.; Cai, S.; Lee, W.C.; et al. The dual functions of the extreme N-terminus of TDP-43 in regulating its biological activity and inclusion formation. Hum. Mol. Genet. 2013, 22, 3112–3122. [Google Scholar] [CrossRef] [PubMed]
  182. Afroz, T.; Hock, E.M.; Ernst, P.; Foglieni, C.; Jambeau, M.; Gilhespy, L.A.B.; Laferriere, F.; Maniecka, Z.; Pluckthun, A.; Mittl, P.; et al. Functional and dynamic polymerization of the ALS-linked protein TDP-43 antagonizes its pathologic aggregation. Nat. Commun. 2017, 8, 45. [Google Scholar] [CrossRef] [PubMed]
  183. Tsoi, P.S.; Choi, K.J.; Leonard, P.G.; Sizovs, A.; Moosa, M.M.; MacKenzie, K.R.; Ferreon, J.C.; Ferreon, A.C.M. The N-Terminal Domain of ALS-Linked TDP-43 Assembles without Misfolding. Angew. Chem. Int. Ed. Engl. 2017, 56, 12590–12593. [Google Scholar] [CrossRef] [PubMed]
  184. Shiina, Y.; Arima, K.; Tabunoki, H.; Satoh, J. TDP-43 dimerizes in human cells in culture. Cell. Mol. Neurobiol. 2010, 30, 641–652. [Google Scholar] [CrossRef] [PubMed]
  185. Fang, Y.S.; Tsai, K.J.; Chang, Y.J.; Kao, P.; Woods, R.; Kuo, P.H.; Wu, C.C.; Liao, J.Y.; Chou, S.C.; Lin, V.; et al. Full-length TDP-43 forms toxic amyloid oligomers that are present in frontotemporal lobar dementia-TDP patients. Nat. Commun. 2014, 5, 4824. [Google Scholar] [CrossRef]
  186. Kayed, R.; Head, E.; Thompson, J.L.; McIntire, T.M.; Milton, S.C.; Cotman, C.W.; Glabe, C.G. Common structure of soluble amyloid oligomers implies common mechanism of pathogenesis. Science 2003, 300, 486–489. [Google Scholar] [CrossRef]
  187. Hyman, A.A.; Weber, C.A.; Julicher, F. Liquid-liquid phase separation in biology. Annu. Rev. Cell Dev. Biol. 2014, 30, 39–58. [Google Scholar] [CrossRef]
  188. Poudyal, R.R.; Pir Cakmak, F.; Keating, C.D.; Bevilacqua, P.C. Physical Principles and Extant Biology Reveal Roles for RNA-Containing Membraneless Compartments in Origins of Life Chemistry. Biochemistry 2018, 57, 2509–2519. [Google Scholar] [CrossRef]
  189. Brangwynne, C.P.; Eckmann, C.R.; Courson, D.S.; Rybarska, A.; Hoege, C.; Gharakhani, J.; Julicher, F.; Hyman, A.A. Germline P granules are liquid droplets that localize by controlled dissolution/condensation. Science 2009, 324, 1729–1732. [Google Scholar] [CrossRef]
  190. Sharma, R.; Choi, K.J.; Quan, M.D.; Sharma, S.; Sankaran, B.; Park, H.; LaGrone, A.; Kim, J.J.; MacKenzie, K.R.; Ferreon, A.C.M.; et al. Liquid condensation of reprogramming factor KLF4 with DNA provides a mechanism for chromatin organization. Nat. Commun. 2021, 12, 5579. [Google Scholar] [CrossRef]
  191. Hernandez-Vega, A.; Braun, M.; Scharrel, L.; Jahnel, M.; Wegmann, S.; Hyman, B.T.; Alberti, S.; Diez, S.; Hyman, A.A. Local Nucleation of Microtubule Bundles through Tubulin Concentration into a Condensed Tau Phase. Cell Rep. 2017, 20, 2304–2312. [Google Scholar] [CrossRef] [Green Version]
  192. Protter, D.S.W.; Parker, R. Principles and Properties of Stress Granules. Trends Cell Biol. 2016, 26, 668–679. [Google Scholar] [CrossRef]
  193. Ambadipudi, S.; Biernat, J.; Riedel, D.; Mandelkow, E.; Zweckstetter, M. Liquid-liquid phase separation of the microtubule-binding repeats of the Alzheimer-related protein Tau. Nat. Commun. 2017, 8, 275. [Google Scholar] [CrossRef]
  194. Sawner, A.S.; Ray, S.; Yadav, P.; Mukherjee, S.; Panigrahi, R.; Poudyal, M.; Patel, K.; Ghosh, D.; Kummerant, E.; Kumar, A.; et al. Modulating alpha-Synuclein Liquid-Liquid Phase Separation. Biochemistry 2021, 60, 3676–3696. [Google Scholar] [CrossRef]
  195. Molliex, A.; Temirov, J.; Lee, J.; Coughlin, M.; Kanagaraj, A.P.; Kim, H.J.; Mittag, T.; Taylor, J.P. Phase separation by low complexity domains promotes stress granule assembly and drives pathological fibrillization. Cell 2015, 163, 123–133. [Google Scholar] [CrossRef]
  196. Kuffner, A.M.; Linsenmeier, M.; Grigolato, F.; Prodan, M.; Zuccarini, R.; Capasso Palmiero, U.; Faltova, L.; Arosio, P. Sequestration within biomolecular condensates inhibits Abeta-42 amyloid formation. Chem. Sci. 2021, 12, 4373–4382. [Google Scholar] [CrossRef]
  197. Cleveland, D.W.; Hwo, S.Y.; Kirschner, M.W. Purification of tau, a microtubule-associated protein that induces assembly of microtubules from purified tubulin. J. Mol. Biol. 1977, 116, 207–225. [Google Scholar] [CrossRef]
  198. Wegmann, S.; Eftekharzadeh, B.; Tepper, K.; Zoltowska, K.M.; Bennett, R.E.; Dujardin, S.; Laskowski, P.R.; MacKenzie, D.; Kamath, T.; Commins, C.; et al. Tau protein liquid-liquid phase separation can initiate tau aggregation. EMBO J. 2018, 37, e98049. [Google Scholar] [CrossRef]
  199. Ferreon, J.; Jain, A.; Choi, K.-J.; Tsoi, P.; MacKenzie, K.; Jung, S.; Ferreon, A. Acetylation Disfavors Tau Phase Separation. Int. J. Mol. Sci. 2018, 19, 1360. [Google Scholar] [CrossRef]
  200. Ray, S.; Singh, N.; Kumar, R.; Patel, K.; Pandey, S.; Datta, D.; Mahato, J.; Panigrahi, R.; Navalkar, A.; Mehra, S.; et al. alpha-Synuclein aggregation nucleates through liquid-liquid phase separation. Nat. Chem. 2020, 12, 705–716. [Google Scholar] [CrossRef]
  201. Babinchak, W.M.; Haider, R.; Dumm, B.K.; Sarkar, P.; Surewicz, K.; Choi, J.K.; Surewicz, W.K. The role of liquid-liquid phase separation in aggregation of the TDP-43 low-complexity domain. J. Biol. Chem. 2019, 294, 6306–6317. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  202. Choi, K.J.; Tsoi, P.S.; Moosa, M.M.; Paulucci-Holthauzen, A.; Liao, S.J.; Ferreon, J.C.; Ferreon, A.C.M. A Chemical Chaperone Decouples TDP-43 Disordered Domain Phase Separation from Fibrillation. Biochemistry 2018, 57, 6822–6826. [Google Scholar] [CrossRef] [PubMed]
  203. Conicella, A.E.; Zerze, G.H.; Mittal, J.; Fawzi, N.L. ALS Mutations Disrupt Phase Separation Mediated by alpha-Helical Structure in the TDP-43 Low-Complexity C-Terminal Domain. Structure 2016, 24, 1537–1549. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Figure 1. Predicted structure and disorder of Aβ, tau, α-syn, and TDP-43. Structures were predicted using AlphaFold [26], and calculations for disorder were determined using IUPRED2 [27,28,29,30]. IUPRED2 provides IUPred and ANCHOR scores for disordered proteins. ANCHOR predicts disorder based on amino acid sequence, identifying potential binding sites that are disordered in isolation, while IUPred predicts disorder based on energy estimation. IUPred scores from 0 to 0.5 are considered ordered while scores from 0.5 to 1 are considered disordered. Aβ-42 (human) was chosen as the representative peptide for Aβ. Tau (human, UniProt ID: P10636) is highly disordered with a C-terminal helical region. α-Syn (human, UniProt ID: P37840) has a disordered C-terminus and α-helical N-terminus. TDP-43 (human, UniProt ID: Q13148) contains a long unstructured C-terminus. (Note that these are predicted structures only and may not represent the major physiological protein conformations. For example, α-syn has been shown to adopt more random coil-like structures in physiological buffer conditions, with the elongated helical structure predicted using AlphaFold only favored in the presence of ligands such as membrane vesicles [31,32,33]).
Figure 1. Predicted structure and disorder of Aβ, tau, α-syn, and TDP-43. Structures were predicted using AlphaFold [26], and calculations for disorder were determined using IUPRED2 [27,28,29,30]. IUPRED2 provides IUPred and ANCHOR scores for disordered proteins. ANCHOR predicts disorder based on amino acid sequence, identifying potential binding sites that are disordered in isolation, while IUPred predicts disorder based on energy estimation. IUPred scores from 0 to 0.5 are considered ordered while scores from 0.5 to 1 are considered disordered. Aβ-42 (human) was chosen as the representative peptide for Aβ. Tau (human, UniProt ID: P10636) is highly disordered with a C-terminal helical region. α-Syn (human, UniProt ID: P37840) has a disordered C-terminus and α-helical N-terminus. TDP-43 (human, UniProt ID: Q13148) contains a long unstructured C-terminus. (Note that these are predicted structures only and may not represent the major physiological protein conformations. For example, α-syn has been shown to adopt more random coil-like structures in physiological buffer conditions, with the elongated helical structure predicted using AlphaFold only favored in the presence of ligands such as membrane vesicles [31,32,33]).
Ijms 24 03380 g001
Figure 2. Molecular underpinnings of ND-linked IDP misfolding and aggregation.
Figure 2. Molecular underpinnings of ND-linked IDP misfolding and aggregation.
Ijms 24 03380 g002
Table 1. Mutations associated with NDs.
Table 1. Mutations associated with NDs.
ProteinMutationCitation
E693G[55,60,61]
* [78]
tauG272V[65,66,79]
P301L[64,65,66,79]
V279M[64]
V337M[64,79]
R406W[65,79]
* [80]
α-SynA53T[40,67,68]
A53E[81]
A30P[67,69]
E46K[67]
* [82]
TDP-43A315T[83]
Q331K[75]
M337V[75]
D169G[83,84]
* [50]
* In-depth review and additional mutations.
Table 2. PTMs associated with NDs.
Table 2. PTMs associated with NDs.
ProteinPTMResidueCitation
PhosphorylationS8[90]
GlycationN-term/K residues[92]
GlycosylationY10[121]
* [122]
tauPhosphorylationS396[97]
S404[97]
S422[97]
AcetylationK280[99]
UbiquitinationK48[103]
K63[105]
* [123]
α-SynPhosphorylationS129[108]
UbiquitinationK6[124]
K12[124]
K23[124]
* [124]
TDP-43PhosphorylationS409[115]
S410[115]
UbiquitinationK48[125]
K63[125]
* [125]
* In-depth review with information about additional PTMs.
Table 3. Truncations associated with NDs.
Table 3. Truncations associated with NDs.
ProteinTruncationCitation
Aβ-38[133]
Aβ-40[133]
Aβ-42[133,134]
* [121]
tautau-314[51,137]
tau-421[51,138]
tau-13[140]
tau-402[140]
tau-45–230[141]
tau-243–441[143]
* [152]
α-Syn1–121[128,148]
* [53]
TDP-43TDP-90–414[149]
TDP-220–414[149,153]
* [154]
* In-depth review and information about additional truncations.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Tsoi, P.S.; Quan, M.D.; Ferreon, J.C.; Ferreon, A.C.M. Aggregation of Disordered Proteins Associated with Neurodegeneration. Int. J. Mol. Sci. 2023, 24, 3380. https://doi.org/10.3390/ijms24043380

AMA Style

Tsoi PS, Quan MD, Ferreon JC, Ferreon ACM. Aggregation of Disordered Proteins Associated with Neurodegeneration. International Journal of Molecular Sciences. 2023; 24(4):3380. https://doi.org/10.3390/ijms24043380

Chicago/Turabian Style

Tsoi, Phoebe S., My Diem Quan, Josephine C. Ferreon, and Allan Chris M. Ferreon. 2023. "Aggregation of Disordered Proteins Associated with Neurodegeneration" International Journal of Molecular Sciences 24, no. 4: 3380. https://doi.org/10.3390/ijms24043380

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

Article Metrics

Back to TopTop