Abstract
The role of microtubule‐associated protein Tau in neurodegeneration has been extensively investigated since the discovery of Tau amyloid aggregates in the brains of patients with Alzheimer's disease (AD). The process of formation of amyloid fibrils is known as amyloidogenesis and attracts much attention as a potential target in the prevention and treatment of neurodegenerative conditions linked to protein aggregation. Cerebral deposition of amyloid aggregates of Tau is observed not only in AD but also in numerous other tauopathies and prion diseases. Amyloidogenesis of intrinsically unstructured monomers of Tau can be triggered by mutations in the Tau gene, post‐translational modifications, or interactions with polyanionic molecules and aggregation‐prone proteins/peptides. The self‐assembly of amyloid fibrils of Tau shares a number of characteristic features with amyloidogenesis of other proteins involved in neurodegenerative diseases. For example, in vitro experiments have demonstrated that the nucleation phase, which is the rate‐limiting stage of Tau amyloidogenesis, is shortened in the presence of fragmented preformed Tau fibrils acting as aggregation templates (“seeds”). Accordingly, Tau aggregates released by tauopathy‐affected neurons can spread the neurodegenerative process in the brain through a prion‐like mechanism, originally described for the pathogenic form of prion protein. Moreover, Tau has been shown to form amyloid strains—structurally diverse self‐propagating aggregates of potentially various pathological effects, resembling in this respect prion strains. Here, we review the current literature on Tau aggregation and discuss mechanisms of propagation of Tau amyloid in the light of the prion‐like paradigm.
Keywords: Tau protein, amyloidogenesis, protein aggregation, tauopathies, Alzheimer's disease
Abbreviations
- Aβ
amyloid β
- AD
Alzheimer's disease
- AEP
asparagine endopeptidase
- AFM
atomic force microscopy
- AGD
argyrophilic grain disease
- ApoE
apolipoprotein E
- APP
amyloid precursor protein
- BSE
bovine spongiform encephalopathy
- CBD
corticobasal degeneration
- Cdk5
cyclin‐dependent kinase 5
- CJD
Creutzfeldt‐Jakob disease
- FAD
familial form of Alzheimer's disease
- FV‐AFM
AFM in the force‐volume mode
- FTDP‐17
frontotemporal dementia and parkinsonism linked to chromosome 17
- GSK3β
glycogen synthase kinase 3β
- GSS
Gerstmann‐Straussler‐Scheinker disease
- HD
Huntington's disease
- IDP
intrinsically disordered protein
- K18
Tau peptide comprising 4 repeats of the microtubule‐binding domain
- K19
Tau peptide comprising 3 repeats of the microtubule‐binding domain
- LTP
long term potentiation
- MAPs
microtubule‐associated proteins
- MAPT
Tau gene
- MARK
microtubule‐associated regulatory kinase
- MBD
microtubule‐binding domain
- MMP‐9
matrix‐metalloproteinase 9
- MTBRs
microtubule‐binding repeats
- MTs
microtubules
- NFTs
neurofibrillary tangles
- NMR
nuclear magnetic resonance
- PHFs
paired helical filaments
- PHF6
a hexapeptide motif comprising a sequence VQIVYK
- PHF6*
a hexapeptide motif comprising a sequence VQIINK
- PHF43
43‐residue peptide corresponding to K19 (265‐338)
- PiD
Pick's disease
- Pin1
peptidylprolyl cis/trans isomerase NIMA‐interacting 1
- PKA
protein kinase A
- PP1
protein phosphatase 1
- PP2A
protein phosphatase 2A
- PP2B
protein phosphatase 2B
- PP5
protein phosphatase 5
- PQC
protein quality control
- PrPC
cellular prion protein
- PrPSc
scrapie isoform of prion protein
- PS1
presenilin 1 gene
- PS2
presenilin 2 gene
- PSP
progressive supranuclear palsy
- PTMs
post‐translational modifications
- SAD
sporadic form of Alzheimer's disease
- SAPK/JNK
stress activated kinase/c‐jun N‐terminal kinase
- SFs
straight filaments
- TEM
transmission electron microscopy
- TGs
transglutaminases
- TSEs
transmissible spongiform encephalopathies
- UPS
ubiquitin–proteasome system.
Introduction
Tau protein, which belongs to the family of microtubule‐associated proteins (MAPs), was discovered in 1975 and identified as a molecule necessary for the assembly and stabilization of microtubular cytoskeleton.1 Subsequently, in Alzheimer's disease (AD), a pathological form of Tau was demonstrated to be the principal component of paired helical filaments (PHFs) which further associate into neurofibrillary tangles (NFTs).2, 3 These structures constitute one of the main neuropathological hallmarks of AD.4 Besides NFTs formation of senile plaques composed of amyloid β (Aβ) peptides is widely recognized as a characteristic trait of AD pathogenesis.5 Deposits of Tau aggregates are also detected in numerous other tauopathies such as frontotemporal dementia and parkinsonism linked to chromosome 17 (FTDP‐17), progressive supranuclear palsy (PSP), corticobasal degeneration (CBD), Pick's disease (PiD), argyrophilic grain disease (AGD), and Down's syndrome (DS).6, 7 Since abundance of tangles in AD brains clearly correlates with the progression of neurodegenerative changes and memory loss, for many years NFTs have been considered toxic.8 However, more recent findings have changed this view, suggesting that certain oligomeric but nonfibrillar and structurally heterogeneous assemblies that appear early and usually transiently during aggregation are the most neurotoxic forms of Tau.9, 10, 11 Amyloid aggregates of Tau, such as PHFs, are formed in a process known as amyloidogenesis or fibrillization. In the course of this process, intrinsically disordered Tau molecules self‐assemble through various intermediates to form highly ordered β‐sheet‐rich amyloid fibrils.12, 13 Typically, PHF aggregates are composed of two fibrils wounding around each other. Tau amyloid assemblies can also adopt several other morphologies such as straight filaments (SFs) and ribbon‐like fibrils.14 The ability of Tau to form polymorphic fibrils parallels similar observations made for other amyloidogenic proteins.15, 16 Amyloid aggregates can propagate their structural features leading to template‐dependent conversion of nonphysiological structures.17, 18 This principle underlies propagation scenarios of mammalian prions which cause transmissible spongiform encephalopathies (TSEs) also known as prion diseases.19 These include inherited, sporadic and acquired diseases, such as kuru, Creutzfeldt‐Jakob disease (CJD), Gerstmann‐Straussler‐Scheinker disease (GSS), scrapie, and bovine spongiform encephalopathy (BSE).19 Pathobiology of prions is linked to the phenomenon of prion strains: transmissible conformational variants of prion protein distinct in terms of pathological patterns. The existence of prion strains has been proposed to explain the observed phenotypic diversity in TSEs.20, 21 Besides prion protein, other amyloidogenic proteins are able to form polymorphic amyloid aggregates capable of self‐propagation to daughter generations of fibrils. It is believed that the underlying molecular mechanisms resemble those implicated in prion replication. The ability to form amyloid strains has been demonstrated for numerous proteins.22, 23, 24, 25 In this regard, it has been proposed that they can be collectively called prion‐like proteins.26, 27 Given the fact that Tau can aggregate via template‐based conversion to form amyloid polymorphs, spreading within specific regions of the brain, it has been proposed that Tau represents an example of the prion‐like protein.22, 28, 29 Nevertheless, the mechanisms of Tau amyloidogenesis, as well as molecular factors which can trigger conformational conversion of unstructured Tau monomers leading to the formation of amyloid assemblies, remain elusive.
Domain Structure of Tau
The microtubule‐associated protein Tau gene (MAPT) is located on chromosome 17q21.30 Alternative splicing of MAPT leads to the generation of six major isoforms of Tau in the adult human brain.31, 32 Splicing of exons 2 and 3, which code for N‐terminal inserts, results in Tau isoforms with two, one or none N‐terminal inserts (2N, 1N, 0N).33 Alternative splicing of exon 10 gives isoforms with four (exon 10 inclusion, 4R) or three microtubule binding repeats (MTBRs) (exon 10 exclusion, 3R).34 The six isoforms range from 352 to 441 amino acid residues in length.35 Each Tau isoform contains a projection domain (encompassing above‐mentioned acidic insertions denoted as N), proline‐rich regions and a microtubule binding domain (MBD) composed of either three (3R) or four (4R) repeats of 31 or 32 residues located in the C‐terminal part of the molecule (Fig. 1).36, 37, 38, 39 The MBD and the proline‐rich regions are both positively charged. The N‐terminal part and a short region at the C‐terminus are acidic. A series of recombinant peptides encompassing the repeat region was constructed including K18 and K19. It turned out that K18 and K19 are prone to aggregation since they do not contain the flanking regions that inhibit amyloidogenesis and they correspond to an amyloid core of PHFs.36 Studies on aggregation of such polypeptides provide important insights into the amyloidogenesis of Tau.
Figure 1.

The domain organization of Tau. (A) There are 6 major Tau isoforms in the human brain: 2N4R, 1N4R, 0N4R, 2N3R, 1N3R, 0N3R. Molecules of these isoforms consist of two main domains: the projection domain that protrudes from the microtubule surface and the microtubule‐binding domain (MBD) which has high affinity for MTs. The isoforms differ in the number of N‐terminal inserts (N) and repeats (R) within microtubule‐binding repeats (MTBRs) region. The MTBRs are located within the MBD. The 4 repeats (4R) isoforms contain two hexapeptide motifs (PHF6* and PHF6) and two cysteines (C291 and C322), whereas 3R isoforms have only one hexapeptide motif (PHF6) and cysteine (C322). The longest isoform (2N4R, 441 amino acid residues) contains two inserts (2N) at the N terminus which are followed by the proline‐rich region (P1, P2) and MTBRs (composed of 4R). The third proline‐rich region (P3) follows the R4. Regions rich in negative (‐) and positive (+) charges are indicated. Tau peptides and polypeptides used as models to study amyloidogenesis include PHF6* and PHF6 (A) as well as K18 (4R), K19 (3R) and PHF43 (B). The predicted structure of Tau polypeptide (267‐312, corresponding to a fragment of the repeat region) in a MT‐bound conformation revealed by NMR spectroscopy is shown in (C) (PDB: 2MZ7,292). Note local enrichment in secondary structures. The structure was visualized in PyMOL293
It has been demonstrated that the C‐terminal fragment of Tau comprising residues 422‐441 adopts α‐helical conformation and inhibits Tau amyloidogenesis (isoform 2N4R) through interaction with residues 321‐375 encompassing the MTBR region.40 In addition, the N‐terminal fragment of Tau, containing residues 1‐196, inhibits Tau fibrillization by interaction with residues 392‐421 at the C‐terminal part, which may lead to the stabilization of Tau monomers.41 Consequently, it has been proposed that association of the C‐terminal part of Tau with the MTBRs prevents the aggregation. Once C‐terminal region moves away from the MTBRs, the N‐terminus may associate with the repeat region, which in turn may allow Tau to adopt an amyloid‐prone conformation.41
Brain‐derived Tau and recombinant Tau polypeptides do not adopt well‐defined structures under physiological conditions.36 The intrinsically disordered character of Tau primarily arises from the excess of positively charged residues and relatively small fraction of bulky hydrophobic amino acid side‐chains.42, 43 In addition, Tau is rich in proline residues which locally prohibit formation of α‐helix or β‐sheet. As a consequence of its amino acid composition, Tau monomers are dynamically interchanging disordered conformations with large solvent‐exposed surfaces.44, 45, 46 The fragment of Tau molecule encompassing the microtubule binding repeats adopts extended conformation as revealed by solution small‐angle X‐ray scattering. Interestingly, the size of the full‐length molecule of Tau is smaller than could be expected, which may be explained by a paperclip/global hairpin structure whose N‐ and C‐termini bend inward in the direction of the repeat region, compensating for its extended conformation.47 Conversely, it has been demonstrated that Tau can adopt a transiently ordered structure upon binding to microtubules which forces the unstructured polypeptide chain of Tau to adopt a biologically functional state with low content of defined secondary structures.48, 49, 50
Tau as an intrinsically disordered protein (IDP) is “sufficiently disordered” for initial stages of aggregation and therefore its aggregation does not face energy barriers related to the partial unfolding which usually precedes amyloidogenesis of structured proteins. In fact, it has been proposed that Tau fibrillization must be associated with folding events which would allow monomers of Tau to acquire partially folded amyloid‐prone structure.51 Furthermore, the large Coulombic charges on each Tau monomer may cause significant obstacles due to electrostatic repulsion which must be overcome or reduced prior to association of the molecules into aggregates.
Molecular Bases of Tau Pathology in Tauopathies
Neurodegenerative diseases linked to protein aggregation are frequently characterized by impairments of the protein quality control (PQC) mechanisms.52 However, it is not clear whether alterations in PQC are primary or secondary events of neurodegeneration. In vitro‐generated or AD‐derived amyloid fibrils of Tau, as well as Tau oligomers can inhibit the ubiquitin–proteasome system (UPS).53, 54 It is also known that distinct forms of Tau: native, carrying mutations and/or aggregated can be degraded by autophagy, which may directly contribute to the reduction of Tau‐linked toxicity.55, 56 Consistently, neurons are sensitive to autophagy impairments which may occur in neurodegenerative diseases such as AD and other tauopathies.57 Molecular chaperones may also prevent amyloidogenesis and loss of their function is frequently observed during neurodegenerative disorders.58 Chaperones have been shown to decrease Tau hyperphosphorylation and aggregation levels, increase the concentration of soluble Tau and support binding of Tau to MTs.59, 60, 61
As mentioned above, tauopathy‐associated aggregates of Tau are structurally‐heterogeneous. Tau can assemble into PHFs, SFs or other linear aggregates.14 In addition, lesion morphology may be diverse, including AD‐typical tangles (NFTs), spindle‐shaped, irregular or diffused deposits.62 Another aspect of the structural heterogeneity of these aggregates manifests on the level of cellular pathology. Pathological changes may affect exclusively neurons, or both neurons and glial cells.62 Furthermore, distinct brain regions are affected by pathological aggregates of Tau.63 Although many factors have been shown to influence Tau amyloidogenesis, it is unclear which of them are primary triggers that impact the progression of tauopathies and precise mechanisms that explain variability among tauopathies remain enigmatic.
According to the amyloid cascade hypothesis,5 increased production and aggregation of Aβ lead to the accumulation of toxic species of this peptide which further causes brain pathology and dementia. This hypothesis is based on the assumption that Tau pathology in AD is a result of Aβ aggregation. The principal role of Aβ in AD is particularly clear in the case of familial form of Alzheimer's disease (FAD). In this form of AD, autosomal dominant mutations in one of three genes: PS1 (presenilin 1), PS2 (presenilin 2), or APP (amyloid precursor protein) are sufficient to cause symptoms of the disease.5 The potential risk factors for sporadic Alzheimer's disease (SAD) include ɛ4 allele of apolipoprotein E (ApoE) which plausibly can influence Aβ clearance mechanisms.64 Importantly, some changes in MAPT expression or alternative splicing have also been suggested to be associated with the risk for AD.65 However, currently there is no described mutation in MAPT involved in AD. It is unclear whether, or to what extent, Tau pathology in AD might be independent of Aβ aggregation. However, it is accepted that AD pathology is inseparably linked to neuronal death mediated by Tau oligomerization/aggregation. Clearly, Tau is essential for toxic effects of Aβ66 since Tau‐depleted hippocampal neurons are not prone to neurodegeneration upon treatment with Aβ amyloid aggregates.67 There seems to be an interplay between Tau and Aβ aggregation—both in vitro and in vivo. For example, Aβ pathology may facilitate Tau pathology in a mouse model of AD by accelerating Tau fibrillization and leading to reduction of density of dendritic spines and cognitive impairment.68 According to some studies, Aβ oligomers can stimulate Tau hyperphosphorylation in hippocampal neurons.69 Furthermore, Aβ dimers have been shown to induce hyperphosphorylation of Tau, impair the MT network, leading to neuritic degeneration in hippocampal neurons.70 Prevention of cognitive impairments in transgenic mice with familial AD mutations and overexpressing hAPP was achieved through the reduction of Tau expression, even at high production of Aβ.71 These findings support the importance of Tau amyloidogenesis in AD and suggest that mutual relationship between Tau and Aβ aggregation may underlie neuropathology in this disease.
In contrast to AD, numerous pathogenic mutations in MAPT gene have been found in the autosomal dominantly inherited dementia FTDP‐17, proving that Tau pathology alone is sufficient to cause neurodegeneration.63 There are over 50 identified mutations in MAPT that are responsible for FTDP‐17, and most of them occur in the microtubule binding domain of Tau.63, 72 It has been demonstrated that mutations in MAPT can stimulate fibrillization of Tau even in the absence of polyanionic inducers.73 Some Tau mutants have increased tendencies to aggregate into PHFs.74 In FTDP‐17, Tau pathology is found in both neurons and glia, varying significantly even among patients with the same mutation in MAPT, which suggests that additional factors may contribute to manifestation of this disorder.7, 75 Interestingly, it has been reported that in FTDP‐17, deposits of Aβ amyloid plaques can also be detected within the brain besides typical aggregates composed of Tau.76 This may imply that amyloid aggregates of Tau can induce Aβ‐related pathology. FTDP‐17 mutations may lead to distinct effects on Tau fibrillization.77 Mutated Tau is hyperphosphorylated in brains of FTDP‐17 patients78 and exhibits a diminished ability to promote assembly of MTs,79 while forming polymorphic amyloid fibrils resembling those found in AD.73 Some mutations in MAPT may lead to aberrant splicing of Tau pre‐mRNA, resulting in the increased ratio of 4R to 3R Tau isoforms, which is approximately 1:1 in the normal adult brain.78, 80, 81
A Model of Tau Amyloidogenesis
According to a widely accepted model, Tau amyloidogenesis proceeds via a nucleated polymerization mechanism which is common for many amyloidogenic proteins. The process of Tau fibrillization displays sigmoidal kinetics and is characterized by three major stages: a lag phase, an elongation phase, and a plateau phase.82, 83 In a process called seeding, upon adding mature amyloid fibrils (seeds) to precursor protein molecules, the elimination of the lag phase and rapid initiation of amyloid growth are observed.84 The recruitment of soluble protein molecules leads to the growth of seeds, which may undergo fragmentation to generate more seeds (Fig. 2).
Figure 2.

The scheme of nucleated formation of Tau amyloid aggregates. (A) Intrinsically unstructured monomers of Tau (indicated in blue) do not acquire one specific conformation, but rather undergo transitions from one accessible conformational state to another. It is assumed that the formation of structured amyloid assemblies from unstructured monomers of Tau requires partial folding. It has been proposed that an amyloid‐competent conformer of Tau (indicated in green) is initially formed. It is also hypothesized that such competent monomers assemble into nuclei of Tau amyloidogenesis. Formation of the nuclei is a rate‐limiting stage and is followed by the rapid growth of fibrils. The tips of the amyloid fibril can act as templates incorporating the competent conformers leading to the elongation of the fibril. Hence, the number of available amyloid ends determines the rate of amyloidogenesis. Fragmentation of amyloid aggregates (usually accomplished in vitro by sonication of mature fibrils) produces short fragments called seeds, which can recruit native and amyloidogenic monomers leading to rapid growth of amyloid assemblies. (B) The scheme of kinetics of unseeded and seeded fibrillization of Tau
Detection of transient entities which appear during the lag phase is difficult and molecular events taking place during this stage remain poorly characterized. As stated earlier, disease‐associated rearrangements within the Tau molecule may lead to association of its amino terminus with MTBRs.41 This conformation has been suggested to promote ordered aggregation of Tau in the presence of inducing agents.41 Formation of amyloid nuclei during the lag phase is slow (due to barriers of mainly entropic nature), but once they are formed, the rapid assembly of amyloid fibrils follows.15, 85, 86 Therefore, formation of nuclei is the principal rate‐limiting step in amyloidogenesis. It has been reported that dimerization of Tau molecules, through intermolecular disulfide bonds, leads to a local antiparallel alignment. This dimerization has been proposed to precede formation of the nucleus.87, 88 Subsequently, it has been speculated that this nucleus consists of 4 to 7 dimers.89 Both Tau dimers with inter‐ and intramolecular disulfide bonds have been proposed to retain ability to form fibrils.90 Conversely, the formation of an intramolecular disulfide bridge between two cysteines in 4R Tau isoforms may lead to the generation of “compact monomers” which were proposed to be fibrillization‐incompetent.91 It should be stressed, however, that the role of dimerization in Tau amyloidogenesis was called into question, regarding very low probability of intermolecular disulfide bonds formation via two cysteines located in the highly positively charged repeat region of 4R Tau molecules.92 Presently, it is clear that formation of disulfide bridges is not crucial for Tau amyloidogenesis to occur. Instead, it has been found that certain short amino acid sequences are particularly important for the initiation of early stages of Tau aggregation. Regions known as hexapeptide motifs, PHF6 (VQIVYK) and PHF6* (VQIINK), may act as nucleating segments for Tau assembly. These sequences are located within the microtubule binding domain (Fig. 1). PHF6 and PHF6* have been proposed to promote aggregation by inducing nucleation events as they have increased tendency to form conformations rich in β‐sheets.73, 93, 94 Soluble protein molecules of Tau are progressively converted into amyloid fibrils during the elongation phase. It has been proposed that the growth of Tau fibrils proceeds through a zipper‐like mechanism in which unfolded molecules of Tau form parallel in‐register β‐strands.95 Successive incorporation of Tau monomers to the tips of fibrils via this mechanism enables growth of the fibril. Finally, amyloidogenesis stops at the plateau phase when maximal amount of protein monomers is assembled into amyloid aggregates.82
Each of the stages of Tau aggregation may be modulated by several factors, including: concentration of precursor protein molecules,96 temperature,97 pH,98 salts,98 shear force,99 as well as additional substances that may either stimulate or inhibit aggregation.100, 101 In vitro studies on Tau fibrillization under nonphysiological or near‐physiological conditions provided important insights into the mechanism of aggregation processes. Conversely, physiologically relevant conditions, under which proteins form amyloid fibrils in the brain, have remained challenging to investigate.102 This can be explained by very low rates of amyloidogenic processes in the cellular environment which evolved to protect native proteins from forming cytotoxic species.103, 104
The Role of Polyanions in Tau Amyloidogenesis
The high abundance of basic amino acid residues within the primary structure of Tau is the main cause of its intrinsically disordered character. This circumstance is expected to have both aggregation‐enhancing (by making Tau monomers more structurally dynamic) and aggregation‐preventing (due to the electrostatic repulsion) consequences. Hence, the excessive positive charge on nascent Tau molecules needs to be effectively attenuated either through an extensive phosphorylation (as is the case of Tau found in NFTs), or through interactions with anionic species (proteinaceous or otherwise). Most studies focusing on the latter scenario look into possible role of physiologically‐relevant poly‐ and oligo‐anions such as glycosaminoglycans (mostly heparin and heparan),105, 106, 107, 108 but also smaller anions with a single negative charge which could be present and involved during in vivo aggregation of Tau—for example arachidonic acid,109 docosahexaenoic acid110 or even taurine.111 It should be noted, however, that in the case of amphiphilic anionic inducers (this category encompasses fatty acids and synthetic detergents—e.g., Ref. 112) the inducer‐Tau association is likely to be preceded by formation of micellar assemblies by the former, i.e., the protein interacts with a macromolecular system with a large net negative charge. There are several studies exploring Tau aggregation under strictly in vitro conditions which employ for this purpose synthetic anionic inducers such as poly‐L‐glutamic acid98, 113 or Congo Red (the latter reviewed in work114). These studies, even if conducted under somehow artificial conditions, may provide important clues into the structural dynamics of aggregation‐prone Tau conformers. In particular, this approach could prove very insightful in regard to certain opaque issues surrounding dynamics of aggregation of Tau which cannot be easily addressed using RNA or heparin, for example the role of chirality and chiral transfer between the inducer and Tau in determining fibrillization pattern. Investigation of mechanisms of Tau aggregation induced by polyglutamic acid may also have important biological implications due to the presence of a polyglutamic sequence located in the C‐terminal region of the tubulin molecule that is involved in binding to the repeat region of Tau.98, 115, 116
However, out of all polyanionic inducers of Tau aggregation, heparin is the most frequently studied and considerably relevant for the physiological conditions, and is therefore the one attracting most attention. Heparin fractions composed of sufficiently long chains are known to be potent polyanionic inducers of Tau fibrillization, which contrasts with the behavior of shorter forms of this glycosaminoglycan.105, 117 Apart from providing compensation of electric charges, heparin chains may act as flexible molecular rulers with periodicity matching that of the protein partner.105, 118 Ramachandran and Udgaonkar119 have shown in their study on four repeat domain of Tau (Tau4RD) that the heparin effect sets in very early in the course of aggregation by promoting binding of two Tau monomers and subsequent formation of an aggregation‐prone dimer. Heparin is particularly useful in studies of aggregation of truncated versions of Tau, and, in general, in vitro heparin‐induced aggregates tend to reveal the proper characteristics of PHFs rather than single filament morphology typically found in arachidonic acid‐induced aggregates (see118 for further discussion). A study by Sibille et al.105 probed with high resolution (afforded by the application of NMR spectroscopy) the precise nature of interactions between full‐length Tau and heparin. The authors have identified major regions within Tau responsible for binding the glycosaminoglycan which include highly charged sequences flanking MTBRs. Furthermore, it was shown that initial interactions with heparin are likely to foster formation of both β‐strands and α‐helices in various locations in Tau which, in the absence of heparin, remain disordered. Finally, the study provided strong evidence that during the aggregation process heparin is buried within the rigid core of growing amyloid fibril.
Further studies are needed to untangle the complex interplay of interactions between Tau and polyanionic molecules during an amyloidogenic pathway and shed light on the possible role of naturally occurring oligoanions in selection of various polymorphs of Tau amyloid.
Post‐Translational Modifications of Tau and Their Role in Amyloidogenesis
The physiological functions of Tau are regulated by various post‐translational modifications (PTMs) which may also affect its conformational state. On the one hand, PTMs may regulate affinity of Tau to MTs.120 On the other hand, an abnormal level of some modifications may trigger Tau aggregation.37 There are contradictory reports on the relevance of some PTMs in Tau amyloidogenesis.121 PTMs of Tau (Table 1) include phosphorylation, proteolytic cleavage, glycosylation, acetylation, prolyl‐isomerization, SUMOylation, ubiquitination, methylation, polyamination, glycation, and nitration.
Table 1.
PTMs of Tau
| PTM | Sites | References |
|---|---|---|
| Phosphorylation | T17, Y18, Y29, T30, T39, S46, T50, T52, S56, S61, T63, S64, S68, T69, T71, T76, T95, T101, T102, T111, S113, T123, S129, S131, T135, S137, T149, T153, T169, T175, T181, S184, S185, S191, S195, Y197, S198, S199, S202, T205, S208, S210, T212, S214, T217, T220, T231, S235, S236, S237, S238, S241, T245, S258, S262, T263, S285, S289, S293, S305, Y310, S316, T319, S320, S324, S341, S352, S356, T361, T373, T377, T386, Y394, S396, S400, T403, S404, S409, S412, S413, T414, S416, S422, T427, S433, S435 | 122, 123, 124, 125, 126 |
| Proteolytic cleavage | D13, N255, D348, N368, E391, D402, D421 | 127, 128, 129, 130 |
| Glycosylation | T181, S199, S202, T205, T212, S214, T217, S262, S356, S404, S422 | 131, 132 |
| Acetylation | K148, K150, K163, K174, K180, K190, K234, K240, K254, K257, K259, K267, K274, K280, K281, K290, K298, K311, K317, K321, K331, K353, K369, K370, K383, K385, K395 | 124, 133, 134, 135, 136 |
| Prolyl‐isomerization | T231 | 137, 138 |
| SUMOylation | K340 | 139 |
| Ubiquitynation | K254, K311, K353 | 124, 140, 141 |
| Methylation | K24, K44, K67, K163, K174, K180, K190, K254, K259, K267, K290, K311, K317, K353, K385 | 142, 143 |
| Polyamination | Q6, K24, Q88, Q124, K163, K174, K180, K190, K225, K234, K240, Q244, Q276, Q288, Q351, K383, K385, Q424 | 131, 144 |
| Glycation | K87, K132, K150, K163, K174, K225, K234, K259, K280, K281, K347, K353, K369 | 145 |
| Nitration | Y18, Y29, Y197, Y310, Y394 | 146, 147 |
Numbering of amino acid residues corresponds to the sequence of the longest human Tau isoform (2N4R, 441 residues). Residues displayed in bold have been found to be modified in AD.
Phosphorylation of Tau
Phosphorylation and dephosphorylation play an important physiological function by modulating affinity of Tau to MTs.148 Hyperphosphorylated Tau accumulates in the form of NFTs in AD and related tauopathies.149, 150, 151 Deposits of hyperphosphorylated Tau are also detected in prion diseases.152 It is widely known that hyperphosphorylated Tau dissociates from MTs which may result in the breakdown of microtubular cytoskeleton.153, 154 Furthermore, in vitro studies have demonstrated that hyperphosphorylated Tau can co‐aggregate with normal Tau leading to the sequestration of the functional molecules, while a treatment of hyperphosphorylated Tau with alkaline phosphatase prevents formation of such co‐aggregates.149 Abnormally elevated phosphorylation compensates for the positive charges on the Tau molecule and thereby may lead to conformational changes of Tau triggering its aggregation.155 A similar effect is observed for above‐mentioned polyanion‐induced amyloidogenesis of Tau.156 Numerous protein kinases which can phosphorylate Tau have been identified, including glycogen synthase kinase 3β (GSK3β),157 protein kinase A (PKA),158 cyclin‐dependent kinase 5 (cdk5),159 microtubule‐associated regulatory kinase (MARK),160 and stress activated kinase/c‐jun N‐terminal kinase (SAPK/JNK).161 Phosphatases which can dephosphorylate Tau in the human brain include PP1, PP2A, PP2B, PP5.162 Among them PP2A is considered as the crucial one.163 It has been proposed that Tau aggregation may be enhanced by the phosphorylation within the C‐terminal part of the molecule, including the microtubule‐binding region, and suppressed when the N‐terminal part and the proline‐rich regions are phosphorylated.130, 164 In tauopathies, Tau becomes hyperphosphorylated at multiple serine and threonine residues.165 Phosphorylation of KXGS motifs, which are located in the microtubule binding domain, precedes Tau amyloidogenesis and promotes dissociation of Tau from MTs.166 Tau in AD‐derived PHFs can be phosphorylated at ∼45 residues.125 Most of phosphorylation sites on Tau molecules are found in the proline‐rich region and the C‐terminal part. Only four phosphorylation sites: S258, S262, S289, S356 are found in the repeat region.125 However, phosphorylation at these residues may significantly affect MT‐binding abilities of Tau molecules. One of the first sites phosphorylated in AD is a residue S262 (located within the microtubule‐binding domain).167 Phosphorylation of numerous other sites has been suggested to influence AD pathogenesis. It remains unclear whether an elevated level of phosphorylation is absolutely required or sufficient to induce Tau aggregation in vivo and to what extent phosphorylated species of Tau are neurotoxic. Furthermore, the phosphorylation at some sites which is thought to cause detachment of Tau from MTs, may also protect Tau from aggregation into PHFs.168 Interestingly, increased level of Tau phosphorylation occurs also in normal fetal brains in the absence of Tau pathology.169 In some animals, Tau is hyperphosphorylated during hibernation as a neuroprotective strategy.170
Proteolytic cleavage of Tau
Proteolytic cleavage can facilitate Tau amyloidogenesis and provide neurotoxic species. The protein is a substrate for several caspases and can be cleaved at residue D421 (isoform 2N4R) in neurons treated with Aβ1‐42 aggregates. It has been shown that Tau 1‐421 assembles during arachidonic acid‐induced fibrillization more readily than the full‐length molecule.129 Caspases, including caspase‐3 and ‐6, have been linked with the proteolytic cleavage of Tau in AD. Caspase‐3 cleaves Tau at residues D25 and D421, whereas caspase‐6 truncates Tau at D13, D402 and D421.171 It has also been demonstrated that activation of calpain‐1 (μ‐calpain) stimulated by Aβ1‐40 aggregates leads to the formation of a 17‐kDa N‐terminal fragment of Tau (Tau 45‐230, 2N4R) which can induce cell death in neurons.172 Another neurotoxic N‐terminal fragment of Tau corresponds to residues 26‐230. An elevated level of this fragment occurs during neuronal apoptosis and is linked to the activation of caspases (calpains may also be involved).173 It has also been found that 26‐28 kDa N‐terminal proteolytic fragments of Tau are present in the CSF of AD patients.174 Conversely, calpain‐1 has been implicated in generating Tau‐CTF24 which is a 24 kDa C‐terminal fragment of Tau cleaved at residue R242.175 Tau‐CTF24 encompasses the repeat region and lacks the N‐terminal part of Tau. This fragment exhibits enhanced susceptibility to undergo heparin‐induced aggregation and does not promote microtubule assembly. An age‐dependent increase in the level of Tau‐CTF24 was found in a mouse model of tauopathy and this fragment is also present in the human brains affected by tauopathies.175 Additionally, some matrix‐metalloproteinases (e.g., MMP‐9) may generate Tau fragments with enhanced propensity to form oligomers which can be internalized by neurons.176 Zhang et al. reported that the proteolytic processing of Tau by asparagine endopeptidase (AEP), a lysosomal cysteine protease, at residues N255 and N368, also leads to the formation of aggregation‐prone fragments which presumably may trigger neurodegeneration.127 Accordingly, the Tau fragment cleaved at N368 has been found in AD patients together with increased activity of AEP. The activity of AEP generates multiple Tau fragments including Tau256‐368 which represents a major part of the repeat region. Tau256‐368 exhibits strong ability to assemble into PHFs during heparin‐induced amyloidogenesis in vitro. Two other AEP‐generated fragments, Tau1‐368 and Tau256‐441, also form typical PHF structures. This contrast with the case of amyloid fibrils formed by Tau1‐255 and Tau369‐441, where PHF morphology has not been observed. Notably, Tau1‐368 and Tau256‐368 have stronger ability to trigger apoptosis in neurons than the other proteolytic fragments, which suggests that the removal of the C‐terminal tail located after N368 residue leads to increased neurotoxicity.127 The mechanisms by which diverse proteolytic fragments of Tau may lead to neurodegenerative changes are unclear. The most amyloidogenic proteolytic fragment of the molecule encompasses the repeat region (“the PHF core”) that has the strong ability to fibrillate and form PHFs.127 Clearly, many proteolytic fragments of Tau are characterized by increased susceptibility to form amyloidogenic and/or toxic intermediates. It has been reported that the truncated forms of Tau participate in the formation of NFTs.129 Proteolytic processing together with hyperphosphorylation of Tau precede formation of NFTs,131 however, it is not clear which of these PTMs may occur first. Nonetheless, in vitro amyloidogenesis of proteolytic fragments of Tau usually requires polyanionic inducers, however some fragments can aggregate even in the absence of polyanions.177 Proteolysis very often leads to the formation of Tau fragments representing the repeat region of Tau (R1‐R4, the residues 244‐368 of the longest Tau isoform). These fragments have increased propensity to aggregate both in vitro and in the brains. As a result, proteolysis which removes fragments flanking the repeat region may be an important mechanism promoting Tau amyloidogenesis. Proteolysis may also generate fragments of Tau whose aggregation is markedly enhanced by the exposition of hexapeptide motifs, PHF6 and PHF6*, which are also located within the repeat region.178 Truncated Tau is found in PHFs. A 12 kDa Tau fragment of approximately 100 residues was identified as the minimal protease‐resistant Tau sequence that represents the core of PHFs in AD.179 Due to the fact that Tau cleaved at E391 (isoform 2N4R) is a component of PHFs in AD, it was proposed that truncation at this site may lead to Tau fibrillization into PHFs and subsequent formation of NFTs.179, 180, 181 To conclude, the proteolytic removal of the sequences flanking MBD provides short aggregation‐prone fragments which can contribute to the progression of Tau pathology in the human brain.
Glycosylation of Tau
A growing body of literature indicates that apart from phosphorylation and proteolytic truncation other PTMs may be involved in Tau amyloidogenesis. Some of these PTMs influence Tau phosphorylation and thereby indirectly modify amyloidogenic properties of the molecule, whereas others may affect stability of PHFs. It has been reported that glycosylation of Tau may lead to opposite effects depending on the type of the modification. Aberrant N‐glycosylation through N‐linkage of glycans at asparagine residues stimulates Tau hyperphosphorylation at several sites catalyzed by some kinases, including PKA, cdk5 and GSK‐3β.123 It has been reported that N‐linked glycans can make Tau molecules more susceptible to hyperphosphorylation and less prone to dephosphorylation.182 It has been observed that deglycosylation of PHFs turns them into bundles composed of single fibrils, therefore it seems that N‐linked glycans may promote formation and stabilization of PHFs by maintaining their helicity.183 Consequently, a possible mechanism explaining the role of N‐glycosylation in Tau amyloidogenesis relies on both stimulation of Tau hyperphosphorylation and stabilization of PHF conformation. Conversely, O‐GlcNAcylation—a type of O‐glycosylation—has been proposed to inhibit Tau phosphorylation which could eventually protect Tau from hyperphosphorylation in AD.184, 185, 186 Glycosylation through O‐linked glycans at S/T sites blocks hydroxyl groups of these residues.123 This prevents their phosphorylation and is likely to inhibit Tau amyloidogenesis. Consistently with that, elevated levels of N‐glycosylation and decreased levels of O‐GlcNAcylation of Tau have been detected in AD brains.123, 187
Acetylation of Tau
Another modification of Tau, namely lysine acetylation precedes formation of NFTs, prevents degradation of hyperphosphorylated Tau and also directly enhances its aggregation in vitro.133, 134 Tau acetylated at K174, K274, K280, and K281 has been found in the brains of patients with AD.133, 135, 136 Furthermore, amyloid aggregates of Tau specifically acetylated at K280, that is located within the PHF6* motif of the repeat region, have been found in the brains of patients with AD and related tauopathies.188, 189 It has been proposed that acetylation of Tau at K280 may lead to dissociation of Tau from MTs.133 Subsequently, acetylated Tau may be released from MTs and becomes prone to other PTMs and fibrillization. Increased levels of acetylation are found in PHFs.133 In a heparin‐induced Tau fibrillization assay, both acetylated monomers of full‐length Tau (2N4R) and K18 show an enhanced ability to form fibrils.133 A possible mechanism that explains the role of Tau acetylation in its amyloidogenesis is based on neutralization of charges within the repeat region, which also decreases the binding of Tau to MTs.134 In particular, hyperphosphorylation may neutralize remaining positive charges on the hyperacetylated Tau.
Prolyl‐isomerization of Tau
Another enzyme which can influence the aggregation of Tau is peptidyl‐prolyl cis/trans‐isomerase (Pin1) catalyzing prolyl‐isomerization. It has been found that Tau phosphorylated at a threonine residue of the T231‐P motif may exist in cis and trans conformations.137 The cis conformer is resistant to dephosphorylation and is prone to aggregation, and is believed to be an early precursor of Tau pathology.190, 191 It has been proposed that Pin1 is involved in the protection against Tau aggregation in AD by isomerizing the phosphorylated molecule from cis to trans conformation.137, 190 Furthermore, Pin1 is required for dephosphorylation of Tau.190
SUMOylation and ubiquitination of Tau
PTMs which are involved in degradation pathways of amyloid aggregates of Tau include SUMOylation and ubiquitination. These two PTMs may display opposite roles in Tau amyloidogenesis. SUMOylation has been shown to inhibit the ubiquitination of Tau, which may contribute to reduced degradation and increased accumulation of aggregated Tau.139 Due to the fact that ubiquitination has been implicated in the targeting of aggregated species of Tau for degradation in the UPS140, 192 its inhibition by Tau SUMOylation may result in enhanced accumulation of Tau amyloid aggregates in neurodegenerative conditions. In AD brains, Tau has been found to be ubiquitinated at three lysine residues: K254, K311, and K353.140, 141 So far, there is evidence for Tau SUMOylation at only one lysine residue—K340.139 Interestingly, SUMOylation stimulates Tau phosphorylation and vice versa.139 It has been proposed that SUMOylation may induce conformational changes in Tau which make it a better substrate for protein kinases.139
Methylation of Tau
Methylation of Tau at lysine residues has been found both in normal and AD brains. The precise role of methylation in Tau biology is, however, unclear. Tau in PHFs derived from AD brains has been found to be methylated at K44, K163, K174, K180, K254, K267, and K290.142 Conversely, Tau methylation also occurs in normal human brains.143 Since methylation very often occurs within MTBRs (seven lysine residues in this region were found to be methylated), it is possible that it may influence affinity of Tau to MTs and affect the amyloidogenesis. Methylation may inhibit Thiazine‐red‐induced fibrillization of Tau by affecting nucleation and elongation phases as well as maturation of Tau fibrils.143 Lysine residues in Tau, which can be methylated, overlap with some sites that can be acetylated, ubiquitinated, as well as glycated and polyaminated (Table 1). Moreover, methylation of Tau may influence its susceptibility for phosphorylation and proteolytic cleavage.142
Polyamination
Activity of transglutaminases (TGs) may lead to incorporation of polyamines (polyamination) into proteins. TGs may also catalyze formation of ε‐(γ‐glutamyl) lysine isopeptide bonds (either inter‐ or intramolecular) involving glutamine and lysine residues.144, 193 Transglutaminase‐catalyzed bonds are found in PHFs and NFTs in AD194 and PSP brains.195 It has been demonstrated that polyaminated Tau has increased stability and is more resistant to calpain‐mediated proteolysis.196 Polyamination has not been shown to affect the ability of Tau to bind MTs196 or stimulate fibrillization.193 The molecular mechanisms by which TGs may contribute to Tau amyloidogenesis are unclear. However, it has been proposed that TGs can link Tau molecules with each other or with some other proteins that are detected in NFTs.144
Nonenzymatic PTMs of Tau
Tau undergoes also nonenzymatic PTMs, including glycation and nitration. It has been reported that glycation can enhance amyloidogenesis of Tau in vitro.197 This PTM does not promote the nucleation phase,197 but can stabilize Tau amyloid fibrils most likely through cross‐linking of Tau molecules.198 Indeed, PHFs are found to be glycated in AD, whereas soluble Tau derived from AD brains is not glycated.198 In vitro, glycation promotes aggregation of AD‐derived Tau into large tangle‐like aggregates formed by PHFs, which correlates with an increased ability of hyperglycated Tau to sediment.198 Glycated recombinant Tau (obtained by incubating Tau with glucose) is more prone to form amyloid fibrils in an inducer‐mediated fibrillization assay than the unmodified protein.197 On the contrary, it has been reported that another nonenzymatic modification—nitration inhibits formation of Tau amyloid fibrils,147 but stimulates Tau oligomerization in vitro.199 Plausibly, nitration may also influence Tau amyloidogenesis in AD and other tauopathies.147 Tau has been shown to be nitrated at five tyrosine residues: Y18, Y29, Y197, Y310, Y394.146, 147 Nitration at Y18, Y29, Y197, and Y394 has been found in AD brains.146, 147 A molecular mechanism which could explain how nitration may affect Tau amyloidogenesis remains unclear.
Characterization of Oligomers and Amyloid Fibrils of Tau
Tau oligomers
As in the case of other amyloidogenic proteins, oligomeric species of Tau are formed transiently during amyloidogenesis and represent highly heterogeneous group of assemblies. It is unclear whether oligomers are on‐ or off‐pathway species. It has been suggested that Tau oligomers may be converted into amyloid fibrils.200 Oligomers are considered to be significantly more toxic than mature fibrils. One of the postulated mechanisms of cytotoxicity of oligomers involves disruption of integrity of biological membranes through permeabilization of lipid bilayers.201 Oligomer‐membrane interactions may lead to the formation of structures that resemble ion channels.202 The oligomers may also interact with other proteins which may lead to the failure of PQC mechanisms and propagation of misfolded proteins within the cell.203 Interestingly, it has been shown that Tau oligomers may appear as a result of traumatic brain injury,204 hyperthermia205 and aging.206 The oligomeric assemblies of Tau have been reported to cause mitochondrial and synaptic dysfunctions,207 inhibit long term potentiation (LTP)208 and disrupt memory in animal models.209 Conversely, oligomeric species of Tau lacking the capacity to cause pathological changes have been reported, as well.210 Tau oligomerization can also be influenced by other amyloidogenic proteins via direct binding. For example, the interaction between Tau273‐284 (a fragment within the second repeat of MTBR encompassing PHF6*) and Aβ25‐35 (an amphipathic fragment of Aβ peptide which retains significant ability to aggregate) may lead to the formation of “heteroligomers” (oligomers made of these two peptides).211
Tau amyloid fibrils
In recent years, numerous studies have discussed the role of amyloid fibrils in the pathology of neurodegenerative diseases, suggesting that they may be less toxic than the oligomers.203, 212, 213 However, there are numerous contradictory reports suggesting that amyloid fibrils of Tau may lead to propagation of amyloid pathology within the brain.17, 22, 214, 215 It seems plausible that Tau fibrils contribute to the pathology that may be initiated by Tau oligomers at the earlier stages of amyloidogenesis.
As mentioned above, the typical forms of Tau amyloid in AD are PHFs (∼95%), whereas SFs are less frequent.216, 217, 218 AD‐derived PHFs are made of all six Tau isoforms.218 These assemblies exhibit helical symmetry, twisted, right handed structure with a periodicity of ∼80 nm and a width of ∼20 nm.217, 219 On the basis of the observation that in vitro‐generated single fibrils of Tau may pair and twist together over incubation period, it has been proposed that SFs may be converted into PHFs.220 Fibrillar species of Tau aggregates have a protease resistant core composed of the microtubule binding repeats.179, 221 It has been proposed that this region is also involved in association of PHFs into NFTs through lateral interactions between filaments which may locally expose their cores.222 This is supported by observation that in vitro‐generated PHFs associate into bundles upon treatment with trypsin, which digests N‐ and C‐terminal parts of Tau amyloid leaving the amyloid core.222 Other proteases can also degrade susceptible components of Tau amyloid leaving the resistant regions.223 Interestingly, PHF morphology has also been reported for in vitro‐generated fibrils assembled from the repeated domain of Tau (K18 peptide) (Fig. 3).
Figure 3.

Amyloid fibrils of Tau visualized by TEM. Recombinant human Tau forms polymorphic fibrils in vitro. These assemblies may resemble amyloid fibrils found in tauopathies. (A) SFs formed by recombinant human full‐length Tau (2N4R, 441 amino acids). (B) PHFs formed by the K18 peptide. The scale bar is 100 nm
On the contrary, SFs represent a minor population of Tau aggregates found in AD and do not exhibit helical morphology and periodicity.218 Hybrid fibrils in which single straight filaments change into PHFs (the transition occurs within a single fibril) can also be found in AD brains.217 These observations lead to the conclusion that amyloid fibrils of Tau do not constitute a homogenous population, but are polymorphic in the brains of AD patients. In FTDP‐17, Tau fibrils made of 3R and/or 4R Tau are predominantly single, but can also appear as PHFs, twisted ribbons and rope‐like fibrils.224 In CBD which is known as a 4R tauopathy,225 Tau aggregates into SFs that are ∼15 nm wide as well as PHFs that are ∼26‐28 nm in their maximal widths and have ∼200 nm intervals.222 In PiD which is a 3R tauopathy,226 Tau assembles into SFs that are ∼15 nm in width and also into PHFs that are ∼26 nm in width and exhibit periodicity of ∼120 nm.227 These findings prove that structure of Tau filaments may differ between AD and non‐AD tauopathies in morphological features. In vitro, the 3R isoforms have tendency to form twisted fibrils, whereas the 4R isoforms preferentially aggregate into single amyloid assemblies.91 It has also been proposed that structural polymorphism of fibrils assembled from 3R and 4R Tau may be influenced by the formation of a disulfide bond (either intermolecular in 3R Tau or intramolecular in 4R Tau).90 It has been demonstrated that in vitro‐generated Tau fibrils share some common characteristics but may also differ in thickness, ability to form twisted or untwisted structures and periodicity.228 Moreover, structural heterogeneity often occurs in a single sample that contains fibrils assembled from identical Tau molecules.
Spectroscopic analysis has brought contradictory data on the content of secondary structures of Tau molecules in the amyloid assemblies. Some groups postulated that Tau amyloid fibrils are β‐sheet rich,229, 230 whereas others suggested lack of cross‐β structure and an unusual α‐helical architecture.231 The characterization of Tau aggregates based on IR‐spectroscopy has not revealed unambiguously the question as to whether amyloid fibrils have structured cores covered with unstructured coats, or fully structured fibrils are mixed with disordered protein molecules. Nevertheless, numerous observations support β‐sheet rich nature of Tau amyloid. Fibrils formed from K18 or K19 polypeptides also have increased content of β‐sheets.73 K18 carrying mutations of FTDP‐17 has been shown to have even more noticeable content of this secondary structure.73 As mentioned, VQIINK and VQIVYK are nucleating segments of Tau that are considered to initiate formation of β‐sheet conformation. VYK (located within the PHF6 motif) is the shortest Tau‐derived peptide able to form amyloid fibrils.232
Full‐length Tau and short polypeptides of Tau representing MTBRs can form fibrils of similar thickness as visualized by TEM and AFM.228 Consequently, it has been proposed that, besides a rigid β‐sheet rich core made of the repeat region, a major part of Tau molecule in fibrils is undetectable by microscopy. This “invisible” component is known as the fuzzy coat. It is a disordered negatively‐charged two‐layered polyelectrolyte brush composed of the N‐terminal and C‐terminal regions of Tau.223 High‐resolution AFM in the force‐volume mode (FV‐AFM) provided evidence that the fuzzy coat is an important integral element of Tau fibrils composed of the full‐length protein.223 The structural characteristics of the fuzzy coat may be influenced by changes in pH and electrolyte concentration.223 Moreover, stability of the fuzzy coat may be regulated by hyperphosphorylation of Tau molecules.233
Cross‐Seeding in Tau Amyloidogenesis
It has been proposed that cross‐seeding between Tau and other amyloidogenic proteins may lead to the emergence of novel amyloid assemblies that exhibit unique properties. In cross‐seeding, preformed amyloid seeds of one protein interact directly with singly dispersed (e.g., in solution) “native” molecules of another yet, sufficiently similar protein causing the latter to convert into amyloid fibrils as well.234, 235, 236 One of the key conditions of cross‐seeding to occur is that the specific conformation of the seed needs to be compatible (i.e., energetically allowed) with the protein being recruited (seeded). In other words, there must be an overlap in the configurational spaces of both proteins corresponding to the same amyloidogenic self‐assembly pattern. Such a pattern (i.e., particular structural variant of amyloid fibril) does not need to be thermodynamically preferred (i.e., most stable) among all amyloid polymorphs accessible to either protein. It is therefore sufficient that molecular docking of monomers at the tip of a structurally‐compatible amyloid seed opens a transition pathway toward amyloid that is faster than de novo aggregation of these monomers, for the cross‐seeding to take place. In other words, the seed added to dispersed monomers provides a transition shortcut to amyloid structure essentially identical, or very similar to the seed itself. This can manifest in the form of so‐called conformational memory effect when daughter fibrils obtained through cross‐seeding follow, with a high fidelity, the biophysical, biochemical, and structural characteristics of the mother seed, rather than the type of amyloid that would be formed through spontaneous (unseeded) aggregation. Hence, cross‐seeding is likely to lead to thermodynamically “frustrated” fibrils which nevertheless remain kinetically stable on the time‐scales of cell life and of typical experiments carried out in vitro. It should be stressed, however, that sheer acceleration of aggregation of one protein by amyloid seeds formed by another (either in vivo or in vitro) may also be a consequence of less specific molecular interactions (e.g., secondary nucleation) which would not warrant the exact structural correspondence between the seed and daughter fibrils. Hence, in principle, we may talk about effective “cross‐seeding” without “conformational memory effect”. Cross‐seeding is either bidirectional (one protein seeds another and vice versa) or unidirectional (one protein seeds another, but it does not happen in the opposite way). However, protein interactions may sometimes lead to the cross‐inhibition of misfolding.234
Cross‐seeding may be implicated in certain mechanisms of pathogenesis in neurodegenerative diseases since co‐deposition of amyloid aggregates of distinct proteins and peptides is frequently observed in these disorders. In particular, both Tau and Aβ amyloid assemblies accumulate in AD brain. It has been observed that Aβ aggregates accelerate Tau fibrillization237 and can force Tau to form PHFs.238 Tau interacts with the central to C‐terminal regions of Aβ peptide.239 The R2 repeat in Tau is involved in direct interaction with Aβ17–42 oligomers.240 Crossbreeding of the mice which express P301L Tau with the mice expressing mutant hAPP leads to enhanced assembly of NFTs, whereas formation of amyloid plaques is not accelerated.241 Furthermore, injection of Aβ1–42 fibrils into the brains of P301L Tau mice enhances formation of NFTs.242 Based on these observations, cross‐seeding between Tau and Aβ has been proposed to be unidirectional.234 It has been reported that Aβ1–42 hexamer protofibrils can bind to K18 and K19. Tau polypeptides cross‐seeded by Aβ may adopt more extended and destabilized conformations which expose their hexapeptide motifs (PHF6 and PHF6*).243 Cross‐seeding between Tau and Aβ has been proposed to proceed through a “stretching‐and‐packing” mechanism.243
Jensen et al. have demonstrated that Tau interacts with C‐terminal part of α‐synuclein molecule through the microtubule‐binding region.244 Subsequently, Tau has been shown to undergo cross‐seeding with α‐synuclein fibrils. Such seeding leads to intracellular accumulation of hyperphosphorylated Tau amyloid aggregates.245 Moreover, distinct strains of α‐synuclein amyloid aggregates may be involved in cross‐seeding of Tau.23 In addition, amyloid aggregates of α‐synuclein have been suggested to be involved in AD pathogenesis.246 It has been proposed that co‐aggregation of Tau and α‐synuclein in neurodegenerative diseases is bidirectional. Interestingly, α‐synuclein has been reported to inhibit Aβ deposition and plaque formation.247 Consequently, understanding the complex relationships between aggregation processes of Tau, α‐synuclein and Aβ in pathogenesis of AD may provide important insights into the molecular mechanisms of this disease.
It has also been postulated that interactions between Tau and huntingtin may influence progression of Huntington's disease (HD). An increased level of expression of 4R Tau isoforms was detected in HD brains in which Tau forms nuclear rod‐like deposits.248 Moreover, Tau is hyperphosphorylated, prone to oligomerization and forms pathological deposits which co‐localize with the aggregates of mutant huntingtin in HD brains.249 Interestingly, expression of HD‐linked mutant of huntingtin has been demonstrated to induce Tau hyperphosphorylation and interaction of intracellular huntingtin inclusions with Tau may lead to the formation of huntingtin/Tau inclusions of ring‐like morphology.250
There are also numerous reports showing that pathological deposits of prion protein aggregates in TSEs may be accompanied by Tau pathology. Aggregates of hyperphosphorylated Tau have been detected in patients with familial and sporadic forms of TSEs,152 as well as in animals experimentally infected with prion diseases.251 Coexistence of prion protein amyloid aggregates and NFTs composed of PHFs was documented in some cases of GSS.252 In addition, Tau was found to be abnormally phosphorylated in CJD.152, 253 Tau interacts both with PrPC and PrPSc.254 Furthermore, some mutants of prion protein exhibit an increased capacity to interact with Tau.255 Due to simultaneous pathology of prion protein and Tau in several TSEs, it is plausible that cross‐seeding may also contribute to the pathomechanisms of prion diseases. It appears likely that cross‐aggregation provides an important source of polymorphic amyloid assemblies which may, for example, exhibit different patterns of spread through distinct brain regions. It remains to be elucidated to what extent cross‐seeding and simultaneous deposition of amyloid aggregates of distinct proteins modulates pathology of tauopathies.
Conformational Memory in Tau Amyloidogenesis
It has been demonstrated that Tau amyloid aggregates isolated from human brains are able to recruit and convert soluble Tau monomers, triggering pathology in the cells and animal models of tauopathies.17, 256 It is not only remarkable that such experimental transmission of tauopathy is possible, but also intriguing that the amyloid structures that emerge upon trans‐organismal seeding acquire conformational traits of the amyloid seeds derived from disease‐affected brains. It is tempting to rationalize this outcome in terms of the conformational memory effect in which daughter amyloid, self‐assembling from the pool of seeded monomers, acquires structural characteristics of the template (seed).257, 258, 259 Such a process of molecular imprinting would parallel other cases of self‐propagation of pathological traits recognized in several neurodegenerative diseases associated with protein misfolding. For example, similar seed‐dependent processes may explain propagation of prions in TSEs. Consequently, mechanisms associated with the template‐based conversion of PrPC into PrPSc may explain the propagation of pathological amyloid aggregates in neurodegenerative disorders.28, 224, 260, 261
As we stated earlier, seed‐controlled proliferation of conformational traits of amyloid aggregates may proceed even if the conditions of aggregation favor alternative amyloid structures. The self‐propagating polymorphism of amyloid fibrils is therefore a generic feature of amyloidogenic proteins25, 258, 262 and should be accessible to Tau, as well. It has been reported that PHFs isolated from AD brains can seed fibrillization of recombinant human 0N4R Tau. Upon such seeding, structural characteristics of PHFs are passed on the nascent amyloid aggregates of recombinant Tau.263 Interestingly, in this study recombinant Tau lacking PTMs (e.g., phosphorylation) was able to adopt conformational features of AD‐derived PHFs. This finding supports the hypothesis that amyloid fibrils formed during neurodegenerative conditions contain the information that can be passed on to native protein molecules. Such prion‐like molecular copying leads to the replication of a pathological structure. The conformational memory effect has also been demonstrated in the case of Tau peptides comprising the repeat region. It has been found that while seeds of K18 efficiently template aggregation of K18 monomers, they are not able to seed aggregation of K19 monomers.264, 265 K19 seeds promote seeded‐aggregation of both K19 and K18 which implies the existence of an asymmetric barrier in seeding between K18 and K19. Surprisingly, fibrils obtained by seeding of K18 monomers with K19 seeds acquire ability to seed K19 even after multiple cycles of seeding, suggesting that K18 is able to form fibrils with specific templating abilities that are able to propagate conformational features.264 However, templating effects in seeded amyloidogenesis of Tau may occur with somehow compromised fidelity. Meyer et al. have reported that amyloid fibrils assembled from K18 molecules are heterogeneous and may convert into another polymorph even in repeated cycles of homologous seeding, whereas K19 forms homogenous fibrils which are not prone to undergo structural changes upon seed‐dependent fibrillization.266 It is possible that the structural changes observed upon seed‐based aggregation of Tau may appear due to, as yet, poorly understood limitations in templating and propagation of Tau amyloid assemblies.267 In principle, changes in the structure of Tau fibrils could be influenced by “structural drift” in which a given conformation of a fibril could evolve into a novel amyloid structure.268 Notably, some kind of “conformational evolution” may also occur during propagation of bona fide prions which have been proposed to exist as an ensemble of several conformations.269, 270, 271 A particular conformer may be prone to propagate in one organism, whereas in the other host its replication might face “the seeding barrier”, favoring amplification of another conformer. Importantly, there are a couple of reasons for phenotypic instability of certain strains. For example, heterogeneity of an amyloid population may reveal itself through a conformational drift observed upon prolonged self‐propagation. An amyloid strain with a kinetic advantage over other amyloid variants will ultimately “win the race” in spreading its structural characteristics. This may change depending on the conditions of the seeding (Fig. 4). There are other possible pathways which could explain conformational drift of amyloid samples involving conformational switching, secondary nucleation and deformed templating.271, 272
Figure 4.

Hypothetical pathways in seeded fibrillization of Tau leading to strains. Tau is able to form structurally distinct amyloid seeds, depending on aggregation conditions. In the upper (A) and middle (B) panels, two types of Tau seeds are presented as single filaments (assembled in condition A indicated in blue) or paired filaments (assembled in condition B indicated in grey). According to the conformational memory hypothesis, these seeds recruit Tau monomers (in green) and convert them into strains which are molecular copies of the maternal seeds even if aggregation environment favors formation of alternative amyloid assemblies. Alternatively, strains may “switch” via conformational transition in which one amyloid seed may induce formation of a strain which may have a different structure. In the bottom (C) panel, two types of seeds are presented in condition D (in brown) that favors assembly of seed D. Conformational drift hypothesis postulates that upon change of aggregation environment, a minor conformer (in this case seed C) may gain a selective advantage, replicating its conformation and dominating the population over time
It has been demonstrated that not all Tau mutants linked to FTDP‐17 can seed wild‐type Tau, implying that distinct Tau seeds have diverse seeding potencies. In particular, full‐length Tau (isoform 0N4R) has been shown to be seeded by amyloid seeds of Tau mutant R406W but not by seeds of P301L mutant, in heparin‐induced fibrillization.273 It is unclear whether such seeding barrier occurs as a result of sequence incompatibility or formation of a particular seeding‐incompetent conformer. In addition, the peptide PHF6 does not possess ability to seed longer Tau peptides.274 Moreover, PHF43 peptide (43‐residue peptide encompassing the third repeat of Tau and flanking residues, N265‐E338, devoid of R2 = V275‐S304) forms single thin untwisted fibrils which can seed aggregation of the shortest isoform of Tau (0N3R). Interestingly, Tau (0N3R) fibrils assembled in the presence of seeds of PHF43 acquire typical PHF morphology.93 Furthermore, fibrillization of Tau 0N3R induced by heparin is manifested by very slow kinetics of aggregation, whereas PHF43 aggregates rapidly. Incubation of Tau 0N3R with amyloid seeds of PHF43 results in significant acceleration of fibrillization.93 Examples mentioned above indicate that the seeded aggregation of Tau resembles the prion‐like process which may, however, exhibit unique features with respect to the templating abilities.
It has been frequently proposed that the progression of Tau pathology is a prion‐like process in which seeded aggregation of normal Tau molecules by pathological seeds leads to the self‐propagation of Tau amyloid aggregates within the brain.17, 22, 28, 275 Distinct phenotypes observed in tauopathies and heterogeneity of these diseases have been linked with the phenomenon resembling prion strains.17, 22 Sanders et al. have demonstrated that HEK293 cells expressing Tau244‐372 accumulate intracellular deposits of Tau aggregates upon treatment with Tau fibrils.22 Injection of lysates of such cells into hippocampi of mice overexpressing Tau mutant P301S caused induction of Tau pathology. Such Tau aggregates can be transferred upon passage through several generations of P301S mice, maintaining their characteristics and the ability to cause brain pathology. Importantly, the inoculation was inefficient in wild‐type mice. It has also been demonstrated that brain homogenates from distinct human tauopathies including AD, AGD, CBD, PiD, and PSP induce specific phenotypes in HEK293 cells.22 Kaufman et al. have isolated 18 distinct Tau amyloid strains from diverse sources.17 Upon treatment of HEK293 cells (expressing the repeat region of 2N4R Tau carrying P301L and V337M mutations) with these variants of Tau amyloids, distinct forms of deposits have been observed (diffused, mosaic, ordered, speckles, disordered, etc). These strains stably propagated and maintained their morphology in the cells during cell division. Moreover, such strains had different seeding activity, and more efficient seeding was associated with stronger toxic effects in the cells. After injection of cell lysates with selected strains into brains of transgenic mice, distinct levels of pathological changes in the brain regions specific for different strains were observed.17 The prion‐like nature of Tau aggregates seems to be supported also by the observation that injection of brain extract from mice overexpressing Tau carrying mutation P301S into brains of mice overexpressing wild‐type Tau leads to the propagation of amyloid pathology within the brain.276 Synthetic Tau amyloid fibrils injected into the brains of mice expressing human P301L Tau induce neurodegenerative changes including hyperphosphorylation and aggregation of Tau, eventually leading to neuronal loss.277
It has been proposed that the spread of Tau aggregation occurs after internalization of Tau amyloid seeds from the extracellular milieu into the cytoplasm most probably through the endocytic pathway.278, 279, 280 The above‐mentioned mechanism may explain typical progress of the pathology observed in tauopathies. It has also been demonstrated that the mice expressing human Tau mutant P301L exclusively in the entorhinal cortex and lacking endogenous (murine) Tau show trans‐synaptic propagation of the mutated Tau in the dentate gyrus in the absence of the endogenous protein, which results in NFTs formation and spread of neurodegeneration.281 This seems to be in contrast to the case of TSE prions where both propagation and neurotoxicity are abolished in PrP‐deficient mice. It has been shown that brain tissue lacking endogenous PrPC is not infected with TSE or damaged after experimental inoculation with bona fide prions.282
Nevertheless, although the prion‐like mechanism of propagation of amyloids is widely accepted, several recent studies have questioned its usefulness with respect to Tau‐related pathologies.267, 281, 283 Differences in the propagation characteristic of Tau amyloid aggregates and bona fide prions can be explained by obvious differences in the biology of Tau and prion protein determined by the primary structure of the molecules. Moreover, in contrary to TSEs, tauopathies do not exhibit infectious etiology and there is no epidemiological data that would suggest transmissibility of these disorders between humans.284, 285, 286 Taking into account these inconsistences, some authors have proposed to classify amyloid aggregates of Tau into a group of “prionoids”.283
Conclusions and Future Directions
Tauopathies remain incurable neurodegenerative diseases in which Tau protein loses its principal function which is the stabilization of microtubular cytoskeleton. Molecular mechanisms of the conformational transitions and association of intrinsically unstructured Tau molecules into highly ordered amyloid assemblies, which are deposited in these diseases, remain challenging to investigate. Nevertheless, several factors have been shown to trigger conformational changes in Tau which may result in amyloidogenic events. These factors include mutations in MAPT, polyanions, several PTMs and cross‐seeding by aggregates composed of other amyloidogenic proteins. The ability to undergo aggregation through seed‐based autocatalytic pathway can liken Tau amyloids to prions. It seems likely that all of amyloid species of Tau can account for the propagation of pathological aggregates. Since Tau may interact with both Aβ and α‐synuclein, studies on cross‐seeding may be of wider importance due to its involvement in neurodegenerative processes. Possibly, cross‐seeding may result in the generation of Tau amyloid polymorphs resembling amyloid strains.
Characterization of processes associated with Tau amyloidogenesis may provide novel targets for therapies that aim to cure tauopathies. Promising approaches have been focusing on targeting misprocessed Tau,287 inhibiting Tau hyperphosphorylation,288, 289 stabilizing MTs by microtubule‐binding drugs,290 blocking Tau seeding and trans‐cellular propagation of Tau aggregates,214, 291 and immunization with Tau oligomer‐specific antibodies.11 Nevertheless, to date, none of these approaches provided effective drugs or therapies. Future studies should focus on better understanding of molecular mechanisms underlying the emergence and the propagation of the amyloid state of Tau as well as the prevention of deleterious interactions between Tau and other amyloidogenic polypeptides.
Acknowledgments
We thank Dr. Hanna Nieznanska for TEM micrographs and the members of the Laboratory of Electron Microscopy at the Nencki Institute of Experimental Biology for expert technical assistance. The micrographs were obtained by means of transmission electron microscope JEM 1400 (JEOL Co., Japan, 2008) equipped with 11 Megapixel TEM Camera MORADA G2 (EMSIS GmbH, Germany) at the Nencki Institute of Experimental Biology. Support from the BST funds of the Faculty of Chemistry, Biological and Chemical Research Centre, University of Warsaw is also gratefully acknowledged.
Grant sponsors: Statutory grant to the Nencki Institute of Experimental Biology from the Ministry of Science and Higher Education. B.N. was supported by scholarship from the College of Inter‐Faculty Individual Studies in Mathematics and Natural Sciences.
References
- 1. Weingarten MD, Lockwood AH, Hwo SY, Kirschner MW (1975) A protein factor essential for microtubule assembly. Proc Natl Acad Sci USA 72:1858–1862. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 2. Kosik KS, Joachim CL, Selkoe DJ (1986) Microtubule‐associated protein tau (tau) is a major antigenic component of paired helical filaments in Alzheimer disease. Proc Natl Acad Sci USA 83:4044–4048. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 3. Wood JG, Mirra SS, Pollock NJ, Binder LI (1986) Neurofibrillary tangles of Alzheimer disease share antigenic determinants with the axonal microtubule‐associated protein tau (tau). Proc Natl Acad Sci USA 83:4040–4043. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 4. Rüb U, Stratmann K, Heinsen H, Seidel K, Bouzrou M, Korf HW (2017) Alzheimer's disease: characterization of the brain sites of the initial Tau cytoskeletal pathology will improve the success of novel immunological anti‐Tau treatment approaches. J Alzheimers Dis 57:683–696. [DOI] [PubMed] [Google Scholar]
- 5. Hardy J, Selkoe DJ (2002) The amyloid hypothesis of Alzheimer's disease: progress and problems on the road to therapeutics. Science 297:353–356. [DOI] [PubMed] [Google Scholar]
- 6. Kovacs GG (2015) Neuropathology of tauopathies: principles and practice. Neuropathol Appl Neurobiol 41:3–23. [DOI] [PubMed] [Google Scholar]
- 7. Hutton M, Lendon CL, Rizzu P, Baker M, Froelich S, Houlden H, Pickering‐Brown S, Chakraverty S, Isaacs A, Grover A, Hackett J, Adamson J, Lincoln S, Dickson D, Davies P, Petersen RC, Stevens M, de Graaff E, Wauters E, van Baren J, Hillebrand M, Joosse M, Kwon JM, Nowotny P, Che LK, Norton J, Morris JC, Reed LA, Trojanowski J, Basun H, Lannfelt L, Neystat M, Fahn S, Dark F, Tannenberg T, Dodd PR, Hayward N, Kwok JB, Schofield PR, Andreadis A, Snowden J, Craufurd D, Neary D, Owen F, Oostra BA, Hardy J, Goate A, van Swieten J, Mann D, Lynch T, Heutink P (1998) Association of missense and 5′‐splice‐site mutations in tau with the inherited dementia FTDP‐17. Nature 393:702–705. [DOI] [PubMed] [Google Scholar]
- 8. Ballatore C, Lee VM‐Y, Trojanowski JQ (2007) Tau‐mediated neurodegeneration in Alzheimer's disease and related disorders. Nat Rev Neurosci 8:663–672. [DOI] [PubMed] [Google Scholar]
- 9. Nilson AN, English KC, Gerson JE, Barton Whittle T, Nicolas Crain C, Xue J, Sengupta U, Castillo‐Carranza DL, Zhang W, Gupta P, Kayed R (2017) Tau oligomers associate with inflammation in the brain and retina of tauopathy mice and in neurodegenerative diseases. J Alzheimers Dis 55:1083–1099. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 10. Laganowsky A, Liu C, Sawaya MR, Whitelegge JP, Park J, Zhao M, Pensalfini A, Soriaga AB, Landau M, Teng PK, Cascio D, Glabe C, Eisenberg D (2012) Atomic view of a toxic amyloid small oligomer. Science 335:1228–1231. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 11. Castillo‐Carranza DL, Sengupta U, Guerrero‐Munoz MJ, Lasagna‐Reeves CA, Gerson JE, Singh G, Estes DM, Barrett ADT, Dineley KT, Jackson GR, Kayed R (2014) Passive immunization with Tau oligomer monoclonal antibody reverses tauopathy phenotypes without affecting hyperphosphorylated neurofibrillary tangles. J Neurosci 34:4260–4272. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 12. Tycko R (2014) Physical and structural basis for polymorphism in amyloid fibrils. Protein Sci 23:1528–1539. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 13. Rochet JC, Lansbury PT (2000) Amyloid fibrillogenesis: themes and variations. Curr Opin Struct Biol 10:60–68. [DOI] [PubMed] [Google Scholar]
- 14. Goedert M, Crowther RA, Spillantini MG (1998) Tau mutations cause frontotemporal dementias. Neuron 21:955–958. [DOI] [PubMed] [Google Scholar]
- 15. Riek R, Eisenberg DS (2016) The activities of amyloids from a structural perspective. Nature 539:227–235. [DOI] [PubMed] [Google Scholar]
- 16. Tycko R (2015) Amyloid polymorphism: structural basis and neurobiological relevance. Neuron 86:632–645. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 17. Kaufman SK, Sanders DW, Thomas TL, Ruchinskas AJ, Vaquer‐Alicea J, Sharma AM, Miller TM, Diamond MI (2016) Tau prion strains dictate patterns of cell pathology, progression rate, and regional vulnerability in vivo. Neuron 92:796–812. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 18. Prusiner SB (2013) Biology and genetics of prion causing neurodegeneration. Annu Rev Genet 47:601–623. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 19. Prusiner SB (1998) Prions. Proc Natl Acad Sci USA 95:13363–13383. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 20. Safar J, Wille H, Itri V, Groth D, Serban H, Torchia M, Cohen FE, Prusiner SB (1998) Eight prion strains have PrP(Sc) molecules with different conformations. Nat Med 4:1157–1165. [DOI] [PubMed] [Google Scholar]
- 21. Tanaka M, Chien P, Naber N, Cooke R, Weissman JS (2004) Conformational variations in an infectious protein determine prion strain differences. Nature 428:323–328. [DOI] [PubMed] [Google Scholar]
- 22. Sanders DW, Kaufman SK, DeVos SL, Sharma AM, Mirbaha H, Li A, Barker SJ, Foley AC, Thorpe JR, Serpell LC, Miller TM, Grinberg LT, Seeley WW, Diamond MI (2014) Distinct tau prion strains propagate in cells and mice and define different tauopathies. Neuron 82:1271–1288. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 23. Guo JL, Covell DJ, Daniels JP, Iba M, Stieber A, Zhang B, Riddle DM, Kwong LK, Xu Y, Trojanowski JQ, Lee VM (2013) Distinct α‐synuclein strains differentially promote Tau inclusions in neurons. Cell 154:103–117. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 24. Watts JC, Condello C, Stohr J, Oehler A, Lee J, DeArmond SJ, Lannfelt L, Ingelsson M, Giles K, Prusiner SB (2014) Serial propagation of distinct strains of Aβ prions from Alzheimer's disease patients. Proc Natl Acad Sci USA 111:10323–10328. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 25. Dzwolak W, Smirnovas V, Jansen R, Winter R (2004) Insulin forms amyloid in a strain‐dependent manner: an FT‐IR spectroscopic study. Protein Sci 13:1927–1932. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 26. Chernoff YO, Lindquist SL, Ono B, Inge‐Vechtomov SG, Liebman SW (1995) Role of the chaperone protein Hsp104 in propagation of the yeast prion‐like factor [psi+]. Science 268:880–884. [DOI] [PubMed] [Google Scholar]
- 27. Frost B, Diamond MI (2010) Prion‐like mechanisms in neurodegenerative diseases. Nat Rev Neurosci 11:155–159. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 28. Jucker M, Walker LC (2013) Self‐propagation of pathogenic protein aggregates in neurodegenerative diseases. Nature 501:45–51. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 29. Furman JL, Vaquer‐Alicea J, White CL, Cairns NJ, Nelson PT, Diamond MI (2017) Widespread tau seeding activity at early Braak stages. Acta Neuropathol 133:91–100. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 30. Poorkaj P, Bird TD, Wijsman E, Nemens E, Garruto RM, Anderson L, Andreadis A, Wiederholt WC, Raskind M, Schellenberg GD (1998) Tau is a candidate gene for chromosome 17 frontotemporal dementia. Ann Neurol 43:815–825. [DOI] [PubMed] [Google Scholar]
- 31. Goedert M, Spillantini MG, Jakes R, Rutherford D, Crowther RA (1989) Multiple isoforms of human microtubule‐associated protein tau: sequences and localization in neurofibrillary tangles of Alzheimer's disease. Neuron 3:519–526. [DOI] [PubMed] [Google Scholar]
- 32. Cox K, Combs B, Abdelmesih B, Morfini G, Brady ST, Kanaan NM (2016) Analysis of isoform‐specific tau aggregates suggests a common toxic mechanism involving similar pathological conformations and axonal transport inhibition. Neurobiol Aging 47:113–126. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 33. Pittman AM, Fung HC, de Silva R (2006) Untangling the tau gene association with neurodegenerative disorders. Hum Mol Genet 15:R188–R195. [DOI] [PubMed] [Google Scholar]
- 34. Trabzuni D, Wray S, Vandrovcova J, Ramasamy A, Walker R, Smith C, Luk C, Gibbs JR, Dillman A, Hernandez DG, Arepalli S, Singleton AB, Cookson MR, Pittman AM, de Silva R, Weale ME, Hardy J, Ryten M (2012) MAPT expression and splicing is differentially regulated by brain region: relation to genotype and implication for tauopathies. Hum Mol Genet 21:4094–4103. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 35. Goedert M, Jakes R (1990) Expression of separate isoforms of human tau protein: correlation with the tau pattern in brain and effects on tubulin polymerization. EMBO J 9:4225–4230. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 36. Mukrasch MD, Biernat J, von Bergen M, Griesinger C, Mandelkow E, Zweckstetter M (2005) Sites of tau important for aggregation populate β‐structure and bind to microtubules and polyanions. J Biol Chem 280:24978–24986. [DOI] [PubMed] [Google Scholar]
- 37. Buée L, Bussière T, Buée‐Scherrer V, Delacourte A, Hof PR (2000) Tau protein isoforms, phosphorylation and role in neurodegenerative disorders. Brain Res Rev 33:95–130. [DOI] [PubMed] [Google Scholar]
- 38. Gustke N, Trinczek B, Biernat J, Mandelkow EM, Mandelkow E (1994) Domains of tau protein and interactions with microtubules. Biochemistry 33:9511–9522. [DOI] [PubMed] [Google Scholar]
- 39. Lee G, Neve RL, Kosik KS (1989) The microtubule binding domain of tau protein. Neuron 2:1615–1624. [DOI] [PubMed] [Google Scholar]
- 40. Berry RW, Abraha A, Lagalwar S, LaPointe N, Gamblin TC, Cryns VL, Binder LI (2003) Inhibition of tau polymerization by its carboxy‐terminal caspase cleavage fragment. Biochemistry 42:8325–8331. [DOI] [PubMed] [Google Scholar]
- 41. Horowitz PM, LaPointe N, Guillozet‐Bongaarts AL, Berry RW, Binder LI (2006) N‐terminal fragments of tau inhibit full‐length tau polymerization in vitro. Biochemistry 45:12859–12866. [DOI] [PubMed] [Google Scholar]
- 42. Cho MK, Kim HY, Bernado P, Fernandez CO, Blackledge M, Zweckstetter M (2007) Amino acid bulkiness defines the local conformations and dynamics of natively unfolded α‐synuclein and tau. J Am Chem Soc 129:3032–3033. [DOI] [PubMed] [Google Scholar]
- 43. Uversky VN (2002) Natively unfolded proteins: a point where biology waits for physics. Protein Sci 11:739–756. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 44. Levine ZA, Larini L, LaPointe NE, Feinstein SC, Shea JE (2015) Regulation and aggregation of intrinsically disordered peptides. Proc Natl Acad Sci USA 112:2758–2763. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 45. van der Lee R, Buljan M, Lang B, Weatheritt RJ, Daughdrill GW, Dunker AK, Fuxreiter M, Gough J, Gsponer J, Jones DT, Kim PM, Kriwacki RW, Oldfield CJ, Pappu RV, Tompa P, Uversky VN, Wright PE, Babu MM (2014) Classification of intrinsically disordered regions and proteins. Chem Rev 114:6589–6631. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 46. Muller‐Spath S, Soranno A, Hirschfeld V, Hofmann H, Ruegger S, Reymond L, Nettels D, Schuler B (2010) Charge interactions can dominate the dimensions of intrinsically disordered proteins. Proc Natl Acad Sci USA 107:14609–14614. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 47. Mylonas E, Hascher A, Bernadó P, Blackledge M, Mandelkow E, Svergun DI (2008) Domain conformation of tau protein studied by solution small‐angle X‐ray scattering. Biochemistry 47:10345–10353. [DOI] [PubMed] [Google Scholar]
- 48. Kadavath H, Hofele RV, Biernat J, Kumar S, Tepper K, Urlaub H, Mandelkow E, Zweckstetter M (2015) Tau stabilizes microtubules by binding at the interface between tubulin heterodimers. Proc Natl Acad Sci USA 112:7501–7506. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 49. Battisti A, Ciasca G, Grottesi A, Bianconi A, Tenenbaum A (2012) Temporary secondary structures in tau, an intrinsically disordered protein. Mol Simul 38:525–533. [Google Scholar]
- 50. Bielska AA, Zondlo NJ (2006) Hyperphosphorylation of tau induces local polyproline II helix. Biochemistry 45:5527–5537. [DOI] [PubMed] [Google Scholar]
- 51. Chirita CN, Congdon EE, Yin H, Kuret J (2005) Triggers of full‐length tau aggregation: a role for partially folded intermediates. Biochemistry 44:5862–5872. [DOI] [PubMed] [Google Scholar]
- 52. Ciechanover A, Kwon YT (2015) Degradation of misfolded proteins in neurodegenerative diseases: therapeutic targets and strategies. Exp Mol Med 47:e147. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 53. Keck S, Nitsch R, Grune T, Ullrich O (2003) Proteasome inhibition by paired helical filament‐tau in brains of patients with Alzheimer's disease. J Neurochem 85:115–122. [DOI] [PubMed] [Google Scholar]
- 54. Ciechanover A, Brundin P (2003) The ubiquitin proteasome system in neurodegenerative diseases: sometimes the chicken, sometimes the egg. Neuron 40:427–446. [DOI] [PubMed] [Google Scholar]
- 55. Krüger U, Wang Y, Kumar S, Mandelkow EM (2012) Autophagic degradation of tau in primary neurons and its enhancement by trehalose. Neurobiol Aging 33:2291–2305. [DOI] [PubMed] [Google Scholar]
- 56. Hamano T, Gendron TF, Causevic E, Yen SH, Lin WL, Isidoro C, Deture M, Ko LW (2008) Autophagic‐lysosomal perturbation enhances tau aggregation in transfectants with induced wild‐type tau expression. Eur J Neurosci 27:1119–1130. [DOI] [PubMed] [Google Scholar]
- 57. Schaeffer V, Goedert M (2012) Stimulation of autophagy is neuroprotective in a mouse model of human tauopathy. Autophagy 8:1686–1687. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 58. Hartl FU, Bracher A, Hayer‐Hartl M (2011) Molecular chaperones in protein folding and proteostasis. Nature 475:324–332. [DOI] [PubMed] [Google Scholar]
- 59. Karagöz GE, Duarte AM, Akoury E, Ippel H, Biernat J, Morán Luengo T, Radli M, Didenko T, Nordhues BA, Veprintsev DB, Dickey CA, Mandelkow E, Zweckstetter M, Boelens R, Madl T, Rüdiger SG (2014) Hsp90‐Tau complex reveals molecular basis for specificity in chaperone action. Cell 156:963–974. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 60. Dou F, Netzer WJ, Tanemura K, Li F, Hartl FU, Takashima A, Gouras GK, Greengard P, Xu H (2003) Chaperones increase association of tau protein with microtubules. Proc Natl Acad Sci USA 100:721–726. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 61. Shimura H, Schwartz D, Gygi SP, Kosik KS (2004) CHIP‐Hsc70 complex ubiquitinates phosphorylated Tau and enhances cell survival. J Biol Chem 279:4869–4876. [DOI] [PubMed] [Google Scholar]
- 62. Ferrer I, López‐González I, Carmona M, Arregui L, Dalfó E, Torrejón‐Escribano B, Diehl R, Kovacs GG (2014) Glial and neuronal tau pathology in tauopathies: characterization of disease‐specific phenotypes and tau pathology progression. J Neuropathol Exp Neurol 73:81–97. [DOI] [PubMed] [Google Scholar]
- 63. Ghetti B, Oblak AL, Boeve BF, Johnson KA, Dickerson BC, Goedert M (2015) Invited review: frontotemporal dementia caused by microtubule‐associated protein tau gene (MAPT) mutations: a chameleon for neuropathology and neuroimaging. Neuropathol Appl Neurobiol 41:24–46. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 64. Castellano JM, Kim J, Stewart FR, Jiang H, DeMattos RB, Patterson BW, Fagan AM, Morris JC, Mawuenyega KG, Cruchaga C, Goate AM, Bales KR, Paul SM, Bateman RJ, Holtzman DM (2011) Human apoE isoforms differentially regulate brain amyloid‐β peptide clearance. Sci Transl Med 3:89ra57. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 65. Jun G, Ibrahim‐Verbaas CA, Vronskaya M, Lambert JC, Chung J, Naj AC, Kunkle BW, Wang LS, Bis JC, Bellenguez C, Harold D, Lunetta KL, Destefano AL, Grenier‐Boley B, Sims R, Beecham GW, Smith AV, Chouraki V, Hamilton‐Nelson KL, Ikram MA, Fievet N, Denning N, Martin ER, Schmidt H, Kamatani Y, Dunstan ML, Valladares O, Laza AR, Zelenika D, Ramirez A, Foroud TM, Choi SH, Boland A, Becker T, Kukull WA, van der Lee SJ, Pasquier F, Cruchaga C, Beekly D, Fitzpatrick AL, Hanon O, Gill M, Barber R, Gudnason V, Campion D, Love S, Bennett DA, Amin N, Berr C, Tsolaki M, Buxbaum JD, Lopez OL, Deramecourt V, Fox NC, Cantwell LB, Tárraga L, Dufouil C, Hardy J, Crane PK, Eiriksdottir G, Hannequin D, Clarke R, Evans D, Mosley TH Jr, Letenneur L, Brayne C, Maier W, De Jager P, Emilsson V, Dartigues JF, Hampel H, Kamboh MI, de Bruijn RF, Tzourio C, Pastor P, Larson EB, Rotter JI, O'Donovan MC, Montine TJ, Nalls MA, Mead S, Reiman EM, Jonsson PV, Holmes C, StGeorge‐Hyslop PH, Boada M, Passmore P, Wendland JR, Schmidt R, Morgan K, Winslow AR, Powell JF, Carasquillo M, Younkin SG, Jakobsdóttir J, Kauwe JS, Wilhelmsen KC, Rujescu D, Nöthen MM, Hofman A, Jones L; IGAP Consortium , Haines JL, Psaty BM, Van Broeckhoven C, Holmans P, Launer LJ, Mayeux R, Lathrop M, Goate AM, Escott‐Price V, Seshadri S, Pericak‐Vance MA, Amouyel P, Williams J, van Duijn CM, Schellenberg GD, Farrer LA, (2016) A novel Alzheimer disease locus located near the gene encoding tau protein. Mol Psychiatry 21:108–117. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 66. Ittner LM, Ke YD, Delerue F, Bi M, Gladbach A, van Eersel J, Wölfing H, Chieng BC, Christie MJ, Napier IA, Eckert A, Staufenbiel M, Hardeman E, Götz J (2010) Dendritic function of tau mediates amyloid‐beta toxicity in Alzheimer's disease mouse models. Cell 142:387–397. [DOI] [PubMed] [Google Scholar]
- 67. Rapoport M, Dawson HN, Binder LI, Vitek MP, Ferreira A (2002) Tau is essential to beta ‐amyloid‐induced neurotoxicity. Proc Natl Acad Sci USA 99:6364–6369. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 68. Chabrier MA, Cheng D, Castello NA, Green KN, LaFerla FM (2014) Synergistic effects of amyloid‐beta and wild‐type human tau on dendritic spine loss in a floxed double transgenic model of Alzheimer's disease. Neurobiol Dis 64:107–117. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 69. De Felice FG, Wu D, Lambert MP, Fernandez SJ, Velasco PT, Lacor PN, Bigio EH, Jerecic J, Acton PJ, Shughrue PJ, Chen‐Dodson E, Kinney GG, Klein WL (2008) Alzheimer's disease‐type neuronal tau hyperphosphorylation induced by Aβ oligomers. Neurobiol Aging 29:1334–1347. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 70. Jin M, Shepardson N, Yang T, Chen G, Walsh D, Selkoe DJ (2011) Soluble amyloid beta‐protein dimers isolated from Alzheimer cortex directly induce Tau hyperphosphorylation and neuritic degeneration. Proc Natl Acad Sci USA 108:5819–5824. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 71. Roberson ED, Scearce‐Levie K, Palop JJ, Yan F, Cheng IH, Wu T, Gerstein H, Yu GQ, Mucke L (2007) Reducing endogenous tau ameliorates amyloid beta‐induced deficits in an Alzheimer's disease mouse model. Science 316:750–754. [DOI] [PubMed] [Google Scholar]
- 72. Orr ME, Sullivan AC, Frost B (2017) A brief overview of tauopathy: causes, consequences, and therapeutic strategies. Trends Pharmacol Sci 38:637–648. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 73. von Bergen M, Barghorn S, Li L, Marx A, Biernat J, Mandelkow EM, Mandelkow E (2001) Mutations of Tau protein in frontotemporal dementia promote aggregation of paired helical filaments by enhancing local beta‐structure. J Biol Chem 276:48165–48174. [DOI] [PubMed] [Google Scholar]
- 74. Nacharaju P, Lewis J, Easson C, Yen S, Hackett J, Hutton M, Yen SH (1999) Accelerated filament formation from tau protein with specific FTDP‐17 missense mutations. FEBS Lett 447:195–199. [DOI] [PubMed] [Google Scholar]
- 75. Wszołek ZK, Slowiński J, Golan M, Dickson DW (2005) Frontotemporal dementia and parkinsonism linked to chromosome 17. Folia Neuropathol 43:258–270. [PubMed] [Google Scholar]
- 76. Ishida C, Kobayashi K, Kitamura T, Ujike H, Iwasa K, Yamada M (2015) Frontotemporal dementia with parkinsonism linked to chromosome 17 with the MAPT R406W mutation presenting with a broad distribution of abundant senile plaques. Neuropathology 35:75–82. [DOI] [PubMed] [Google Scholar]
- 77. Combs B, Gamblin TC (2012) FTDP‐17 tau mutations induce distinct effects on aggregation and microtubule interactions. Biochemistry 51:8597–8607. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 78. Goedert M, Spillantini MG (2000) Tau mutations in frontotemporal dementia FTDP‐17 and their relevance for Alzheimer's disease. Biochim Biophys Acta 1502:110–121. [DOI] [PubMed] [Google Scholar]
- 79. Hasegawa M, Smith MJ, Goedert M (1998) Tau proteins with FTDP‐17 mutations have a reduced ability to promote microtubule assembly. FEBS Lett 437:207–210. [DOI] [PubMed] [Google Scholar]
- 80. Goedert M, Jakes R (2005) Mutations causing neurodegenerative tauopathies. Biochim Biophys Acta 1739:240–250. [DOI] [PubMed] [Google Scholar]
- 81. McCarthy A, Lonergan R, Olszewska DA, O'Dowd S, Cummins G, Magennis B, Fallon EM, Pender N, Huey ED, Cosentino S, O'Rourke K, Kelly BD, O'Connell M, Delon I, Farrell M, Spillantini MG, Rowland LP, Fahn S, Craig P, Hutton M, Lynch T (2015) Closing the tau loop: the missing tau mutation. Brain 138:3100–3109. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 82. Lee CC, Nayak A, Sethuraman A, Belfort G, McRae GJ (2007) A three‐stage kinetic model of amyloid fibrillation. Biophys J 92:3448–3458. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 83. Kuret J, Chirita CN, Congdon EE, Kannanayakal T, Li G, Necula M, Yin H, Zhong Q (2005) Pathways of tau fibrillization. Biochim Biophys Acta 1739:167–178. [DOI] [PubMed] [Google Scholar]
- 84. Colby DW, Zhang Q, Wang S, Groth D, Legname G, Riesner D, Prusiner SB (2007) Prion detection by an amyloid seeding assay. Proc Natl Acad Sci USA 104:20914–20919. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 85. Ghosh P, Vaidya A, Kumar A, Rangachari V (2016) Determination of critical nucleation number for a single nucleation amyloid‐β aggregation model. Math Biosci 273:70–79. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 86. Langkilde AE, Vestergaard B (2009) Methods for structural characterization of prefibrillar intermediates and amyloid fibrils. FEBS Lett 583:2600–2609. [DOI] [PubMed] [Google Scholar]
- 87. Wille H, Drewes G, Biernat J, Mandelkow EM, Mandelkow E (1992) Alzheimer‐like paired helical filaments and antiparallel dimers formed from microtubule‐associated protein tau in vitro. J Cell Biol 118:573–584. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 88. Schweers O, Mandelkow EM, Biernat J, Mandelkow E (1995) Oxidation of cysteine‐322 in the repeat domain of microtubule‐associated protein tau controls the in vitro assembly of paired helical filaments. Proc Natl Acad Sci USA 92:8463–8467. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 89. Friedhoff P, von Bergen M, Mandelkow EM, Davies P, Mandelkow E (1998) A nucleated assembly mechanism of Alzheimer paired helical filaments. Proc Natl Acad Sci USA 95:15712–15717. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 90. Furukawa Y, Kaneko K, Nukina N (2011) Tau protein assembles into isoform‐ and disulfide‐dependent polymorphic fibrils with distinct structural properties. J Biol Chem 286:27236–27246. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 91. Barghorn S, Mandelkow E (2002) Toward a unified scheme for the aggregation of tau into Alzheimer paired helical filaments. Biochemistry 41:14885–14896. [DOI] [PubMed] [Google Scholar]
- 92. Rosenberg KJ, Ross JL, Feinstein HE, Feinstein SC, Israelachvili J (2008) Complementary dimerization of microtubule‐associated tau protein: implications for microtubule bundling and tau‐mediated pathogenesis. Proc Natl Acad Sci USA 105:7445–7450. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 93. von Bergen M, Friedhoff P, Biernat J, Heberle J, Mandelkow EM, Mandelkow E (2000) Assembly of tau protein into Alzheimer paired helical filaments depends on a local sequence motif ((306)VQIVYK(311)) forming beta structure. Proc Natl Acad Sci USA 97:5129–5134. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 94. Ganguly P, Do TD, Larini L, LaPointe NE, Sercel AJ, Shade MF, Feinstein SC, Bowers MT, Shea JE (2015) Tau assembly: the dominant role of PHF6 (VQIVYK) in microtubule binding region repeat R3. J Phys Chem B 119:4582–4593. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 95. Margittai M, Langen R (2004) Template‐assisted filament growth by parallel stacking of tau. Proc Natl Acad Sci U S A 101:10278–10283. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 96. Powers ET, Powers DL (2006) The kinetics of nucleated polymerizations at high concentrations: amyloid fibril formation near and above the “supercritical concentration.” Biophys J 91:122–132. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 97. Luo Y, Dinkel P, Yu X, Margittai M, Zheng J, Nussinov R, Wei G, Ma B (2013) Molecular insights into the reversible formation of tau protein fibrils. Chem Commun 49:3582–3584. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 98. Friedhoff P, Schneider A, Mandelkow EM, Mandelkow E (1998) Rapid assembly of Alzheimer‐like paired helical filaments from microtubule‐associated protein tau monitored by fluorescence in solution. Biochemistry 37:10223–10230. [DOI] [PubMed] [Google Scholar]
- 99. Adamcik J, Sánchez‐Ferrer A, Ait‐Bouziad N, Reynolds NP, Lashuel HA, Mezzenga R (2016) Microtubule‐binding R3 fragment from Tau self‐assembles into giant multistranded amyloid ribbons. Angew Chem Int Ed Engl 55:618–622. [DOI] [PubMed] [Google Scholar]
- 100. Dinkel PD, Holden MR, Matin N, Margittai M (2015) RNA binds to Tau fibrils and sustains template‐assisted growth. Biochemistry 54:4731–4740. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 101. Paranjape SR, Chiang YM, Sanchez JF, Entwistle R, Wang CC, Oakley BR, Gamblin TC (2014) Inhibition of tau aggregation by three Aspergillus nidulans secondary metabolites: 2, ω ‐dihydroxyemodin, asperthecin, and asperbenzaldehyde. Planta Med 80:77–85. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 102. Jahn TR, Parker MJ, Homans SW, Radford SE (2006) Amyloid formation under physiological conditions proceeds via a native‐like folding intermediate. Nat Struct Mol Biol 13:195–201. [DOI] [PubMed] [Google Scholar]
- 103. Monsellier E, Chiti F (2007) Prevention of amyloid‐like aggregation as a driving force of protein evolution. EMBO Rep 8:737–742. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 104. Lansbury PT (1999) Evolution of amyloid: what normal protein folding may tell us about fibrillogenesis and disease. Proc Natl Acad Sci USA 96:3342–3344. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 105. Sibille N, Sillen A, Leroy A, Wieruszeski JM, Mulloy B, Landrieu I, Lippens G (2006) Structural impact of heparin binding to full‐length Tau as studied by NMR spectroscopy. Biochemistry 45:12560–12572. [DOI] [PubMed] [Google Scholar]
- 106. Goedert M, Jakes R, Spillantini MG, Hasegawa M, Smith MJ, Crowther RA (1996) Assembly of microtubule‐associated protein tau into Alzheimer‐like filaments induced by sulphated glycosaminoglycans. Nature 383:550–553. [DOI] [PubMed] [Google Scholar]
- 107. Holmes BB, DeVos SL, Kfoury N, Li M, Jacks R, Yanamandra K, Ouidja MO, Brodsky FM, Marasa J, Bagchi DP, Kotzbauer PT, Miller TM, Papy‐Garcia D, Diamond MI (2013) Heparan sulfate proteoglycans mediate internalization and propagation of specific proteopathic seeds. Proc Natl Acad Sci USA 110:E3138–E3147. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 108. Kampers T, Friedhoff P, Biernat J, Mandelkow EM, Mandelkow E (1996) RNA stimulates aggregation of microtubule‐associated protein tau into Alzheimer‐like paired helical filaments. FEBS Lett 399:344–349. [DOI] [PubMed] [Google Scholar]
- 109. Wilson DM, Binder LI (1997) Free fatty acids stimulate the polymerization of tau and amyloid beta peptides. In vitro evidence for a common effector of pathogenesis in Alzheimer's disease. Am J Pathol 150:2181–2195. [PMC free article] [PubMed] [Google Scholar]
- 110. Gamblin TC, King ME, Kuret J, Berry RW, Binder LI (2000) Oxidative regulation of fatty acid‐induced tau polymerization. Biochemistry 39:14203–14210. [DOI] [PubMed] [Google Scholar]
- 111. Santa‐María I, Hernández F, Moreno FJ, Avila J (2007) Taurine, an inducer for tau polymerization and a weak inhibitor for amyloid‐beta‐peptide aggregation. Neurosci Lett 429:91–94. [DOI] [PubMed] [Google Scholar]
- 112. Chirita CN, Necula M, Kuret J (2003) Anionic micelles and vesicles induce tau fibrillization in vitro. J Biol Chem 278:25644–25650. [DOI] [PubMed] [Google Scholar]
- 113. Akoury E, Mukrasch MD, Biernat J, Tepper K, Ozenne V, Mandelkow E, Blackledge M, Zweckstetter M (2016) Remodeling of the conformational ensemble of the repeat domain of tau by an aggregation enhancer. Protein Sci 25:1010–1020. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 114. Chang E, Honson NS, Bandyopadhyay B, Funk KE, Jensen JR, Kim S, Naphade S, Kuret J (2009) Modulation and detection of tau aggregation with small‐molecule ligands. Curr Alzheimer Res 6:409–414. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 115. Avila J, Pérez M, Lucas JJ, Gómez‐Ramos A, Santa María I, Moreno F, Smith M, Perry G, Hernández F (2004) Assembly in vitro of tau protein and its implications in Alzheimer's disease. Curr Alzheimer Res 1:97–101. [DOI] [PubMed] [Google Scholar]
- 116. Eddé B, Rossier J, Le Caer JP, Desbruyères E, Gros F, Denoulet P (1990) Posttranslational glutamylation of alpha‐tubulin. Science 247:83–85. [DOI] [PubMed] [Google Scholar]
- 117. Jangholi A, Ashrafi‐Kooshk MR, Arab SS, Riazi G, Mokhtari F, Poorebrahim M, Mahdiuni H, Kurganov BI, Moosavi‐Movahedi AA, Khodarahmi R (2016) Appraisal of role of the polyanionic inducer length on amyloid formation by 412‐residue 1N4R Tau protein: a comparative study. Arch Biochem Biophys 609:1–19. [DOI] [PubMed] [Google Scholar]
- 118. Gamblin TC, Berry RW, Binder LI (2003) Modeling tau polymerization in vitro: a review and synthesis. Biochemistry 42:15009–15017. [DOI] [PubMed] [Google Scholar]
- 119. Ramachandran G, Udgaonkar JB (2011) Understanding the kinetic roles of the inducer heparin and of rod‐like protofibrils during amyloid fibril formation by Tau protein. J Biol Chem 286:38948–38959. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 120. Mandelkow EM, Mandelkow E (2012) Biochemistry and cell biology of tau protein in neurofibrillary degeneration. Cold Spring Harb Perspect Med 2:a006247. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 121. Lee HG, Perry G, Moreira PI, Garrett MR, Liu Q, Zhu X, Takeda A, Nunomura A, Smith MA (2005) Tau phosphorylation in Alzheimer's disease: pathogen or protector? Trends Mol Med 11:164–169. [DOI] [PubMed] [Google Scholar]
- 122. Beharry C, Cohen LS, Di J, Ibrahim K, Briffa‐Mirabella S, Alonso AC (2014) Tau‐induced neurodegeneration: mechanisms and targets. Neurosci Bull 30:346–358. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 123. Liu F, Iqbal K, Grundke‐Iqbal I, Gong CX (2002) Involvement of aberrant glycosylation in phosphorylation of tau by cdk5 and GSK‐3beta. FEBS Lett 530:209–214. [DOI] [PubMed] [Google Scholar]
- 124. Cook C, Stankowski JN, Carlomagno Y, Stetler C, Petrucelli L (2014) Acetylation: a new key to unlock tau's role in neurodegeneration. Alzheimers Res Ther 6:29. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 125. Hanger DP, Anderton BH, Noble W (2009) Tau phosphorylation: the therapeutic challenge for neurodegenerative disease. Trends Mol Med 15:112–119. [DOI] [PubMed] [Google Scholar]
- 126. Hanger DP, Byers HL, Wray S, Leung KY, Saxton MJ, Seereeram A, Reynolds CH, Ward MA, Anderton BH (2007) Novel phosphorylation sites in tau from Alzheimer brain support a role for casein kinase 1 in disease pathogenesis. J Biol Chem 282:23645–23654. [DOI] [PubMed] [Google Scholar]
- 127. Zhang Z, Song M, Liu X, Kang SS, Kwon IS, Duong DM, Seyfried NT, Hu WT, Liu Z, Wang JZ, Cheng L, Sun YE, Yu SP, Levey AI, Ye K (2014) Cleavage of tau by asparagine endopeptidase mediates the neurofibrillary pathology in Alzheimer's disease. Nat Med 20:1254–1262. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 128. Pritchard SM, Dolan PJ, Vitkus A, Johnson GV (2011) The toxicity of tau in Alzheimer disease: turnover, targets and potential therapeutics. J Cell Mol Med 15:1621–1635. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 129. Gamblin TC, Chen F, Zambrano A, Abraha A, Lagalwar S, Guillozet AL, Lu M, Fu Y, Garcia‐Sierra F, LaPointe N, Miller R, Berry RW, Binder LI, Cryns VL (2003) Caspase cleavage of tau: linking amyloid and neurofibrillary tangles in Alzheimer's disease. Proc Natl Acad Sci USA 100:10032–10037. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 130. Fontaine SN, Sabbagh JJ, Baker J, Martinez‐Licha CR, Darling A, Dickey CA (2015) Cellular factors modulating the mechanism of tau protein aggregation. Cell Mol Life Sci 72:1863–1879. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 131. Martin L, Latypova X, Terro F (2011) Post‐translational modifications of tau protein: implications for Alzheimer's disease. Neurochem Int 58:458–471. [DOI] [PubMed] [Google Scholar]
- 132. Yu CH, Si T, Wu WH, Hu J, Du JT, Zhao YF, Li YM (2008) O‐GlcNAcylation modulates the self‐aggregation ability of the fourth microtubule‐binding repeat of tau. Biochem Biophys Res Commun 375:59–62. [DOI] [PubMed] [Google Scholar]
- 133. Cohen TJ, Guo JL, Hurtado DE, Kwong LK, Mills IP, Trojanowski JQ, Lee VM (2011) The acetylation of tau inhibits its function and promotes pathological tau aggregation. Nat Commun 2:252. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 134. Min SW, Cho SH, Zhou Y, Schroeder S, Haroutunian V, Seeley WW, Huang EJ, Shen Y, Masliah E, Mukherjee C, Meyers D, Cole PA, Ott M, Gan L (2010) Acetylation of tau inhibits its degradation and contributes to tauopathy. Neuron 67:953–966. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 135. Min SW, Chen X, Tracy TE, Li Y, Zhou Y, Wang C, Shirakawa K, Minami SS, Defensor E, Mok SA, Sohn PD, Schilling B, Cong X, Ellerby L, Gibson BW, Johnson J, Krogan N, Shamloo M, Gestwicki J, Masliah E, Verdin E, Gan L (2015) Critical role of acetylation in tau‐mediated neurodegeneration and cognitive deficits. Nat Med 21:1154–1162. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 136. Tracy TE, Sohn PD, Minami SS, Wang C, Min SW, Li Y, Zhou Y, Le D, Lo I, Ponnusamy R, Cong X, Schilling B, Ellerby LM, Huganir RL, Gan L (2016) Acetylated Tau obstructs KIBRA‐mediated signaling in synaptic plasticity and promotes tauopathy‐related memory loss. Neuron 90:245–260. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 137. Lu PJ, Wulf G, Zhou XZ, Davies P, Lu KP (1999) The prolyl isomerase Pin1 restores the function of Alzheimer‐associated phosphorylated tau protein. Nature 399:784–788. [DOI] [PubMed] [Google Scholar]
- 138. Nakamura K, Greenwood A, Binder L, Bigio EH, Denial S, Nicholson L, Zhou XZ, Lu KP (2012) Proline isomer‐specific antibodies reveal the early pathogenic tau conformation in Alzheimer's disease. Cell 149:232–244. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 139. Luo HB, Xia YY, Shu XJ, Liu ZC, Feng Y, Liu XH, Yu G, Yin G, Xiong YS, Zeng K, Jiang J, Ye K, Wang XC, Wang JZ (2014) SUMOylation at K340 inhibits tau degradation through deregulating its phosphorylation and ubiquitination. Proc Natl Acad Sci USA 111:16586–16591. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 140. Cripps D, Thomas SN, Jeng Y, Yang F, Davies P, Yang AJ (2006) Alzheimer disease‐specific conformation of hyperphosphorylated paired helical filament‐Tau is polyubiquitinated through Lys‐48, Lys‐11, and Lys‐6 ubiquitin conjugation. J Biol Chem 281:10825–10838. [DOI] [PubMed] [Google Scholar]
- 141. Thomas SN, Cripps D, Yang AJ (2009) Proteomic analysis of protein phosphorylation and ubiquitination in Alzheimer's disease. Methods Mol Biol 566:109–121. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 142. Thomas SN, Funk KE, Wan Y, Liao Z, Davies P, Kuret J, Yang AJ (2012) Dual modification of Alzheimer's disease PHF‐tau protein by lysine methylation and ubiquitylation: a mass spectrometry approach. Acta Neuropathol 123:105–117. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 143. Funk KE, Thomas SN, Schafer KN, Cooper GL, Liao Z, Clark DJ, Yang AJ, Kuret J (2014) Lysine methylation is an endogenous post‐translational modification of tau protein in human brain and a modulator of aggregation propensity. Biochem J 462:77–88. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 144. Murthy SN, Wilson JH, Lukas TJ, Kuret J, Lorand L (1998) Cross‐linking sites of the human tau protein, probed by reactions with human transglutaminase. J Neurochem 71:2607–2614. [DOI] [PubMed] [Google Scholar]
- 145. Nacharaju P, Ko L, Yen SH (1997) Characterization of in vitro glycation sites of tau. J Neurochem 69:1709–1719. [DOI] [PubMed] [Google Scholar]
- 146. Reyes JF, Fu Y, Vana L, Kanaan NM, Binder LI (2011) Tyrosine nitration within the proline‐rich region of tau in Alzheimer's disease. Am J Pathol 178:2275–2285. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 147. Reynolds MR, Reyes JF, Fu Y, Bigio EH, Guillozet‐Bongaarts AL, Berry RW, Binder LI (2006) Tau nitration occurs at tyrosine 29 in the fibrillar lesions of Alzheimer's disease and other tauopathies. J Neurosci 26:10636–10645. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 148. Lindwall G, Cole RD (1984) Phosphorylation affects the ability of tau protein to promote microtubule assembly. J Biol Chem 259:5301–5305. [PubMed] [Google Scholar]
- 149. Alonso AC, Grundke‐Iqbal I, Iqbal K (1996) Alzheimer's disease hyperphosphorylated tau sequesters normal tau into tangles of filaments and disassembles microtubules. Nat Med 2:783–787. [DOI] [PubMed] [Google Scholar]
- 150. Alonso AC, Mederlyova A, Novak M, Grundke‐Iqbal I, Iqbal K (2004) Promotion of hyperphosphorylation by frontotemporal dementia tau mutations. J Biol Chem 279:34873–34881. [DOI] [PubMed] [Google Scholar]
- 151. Iqbal K, Alonso Adel C, Chen S, Chohan MO, El‐Akkad E, Gong CX, Khatoon S, Li B, Liu F, Rahman A, Tanimukai H, Grundke‐Iqbal I (2005) Tau pathology in Alzheimer disease and other tauopathies. Biochim Biophys Acta 1739:198–210. [DOI] [PubMed] [Google Scholar]
- 152. Sikorska B, Liberski PP, Sobów T, Budka H, Ironside JW (2009) Ultrastructural study of florid plaques in variant Creutzfeldt‐Jakob disease: a comparison with amyloid plaques in kuru, sporadic Creutzfeldt‐Jakob disease and Gerstmann‐Sträussler‐Scheinker disease. Neuropathol Appl Neurobiol 35:46–59. [DOI] [PubMed] [Google Scholar]
- 153. Alonso AC, Zaidi T, Grundke‐Iqbal I, Iqbal K (1994) Role of abnormally phosphorylated tau in the breakdown of microtubules in Alzheimer disease. Proc Natl Acad Sci USA 91:5562–5566. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 154. Iqbal K, Liu F, Gong CX (2016) Tau and neurodegenerative disease: the story so far. Nat Rev Neurol 12:15–27. [DOI] [PubMed] [Google Scholar]
- 155. Alonso AD, Zaidi T, Novak M, Barra HS, Grundke‐Iqbal I, Iqbal K (2001) Interaction of tau isoforms with Alzheimer's disease abnormally hyperphosphorylated tau and in vitro phosphorylation into the disease‐like protein. J Biol Chem 276:37967–37973. [DOI] [PubMed] [Google Scholar]
- 156. Jeganathan S, von Bergen M, Mandelkow EM, Mandelkow E (2008) The natively unfolded character of Tau and its aggregation to Alzheimer‐like paired helical filaments. Biochemistry 47:10526–10539. [DOI] [PubMed] [Google Scholar]
- 157. Lucas JJ, Hernández F, Gómez‐Ramos P, Morán MA, Hen R, Avila J (2001) Decreased nuclear beta‐catenin, tau hyperphosphorylation and neurodegeneration in GSK‐3beta conditional transgenic mice. EMBO J 20:27–39. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 158. Liu SJ, Zhang JY, Li HL, Fang ZY, Wang Q, Deng HM, Gong CX, Grundke‐Iqbal I, Iqbal K, Wang JZ (2004) Tau becomes a more favorable substrate for GSK‐3 when it is prephosphorylated by PKA in rat brain. J Biol Chem 279:50078–50088. [DOI] [PubMed] [Google Scholar]
- 159. Yamaguchi H, Ishiguro K, Uchida T, Takashima A, Lemere CA, Imahori K (1996) Preferential labeling of Alzheimer neurofibrillary tangles with antisera for tau protein kinase (TPK) I/glycogen synthase kinase‐3 beta and cyclin‐dependent kinase 5, a component of TPK II. Acta Neuropathol 92:232–241. [DOI] [PubMed] [Google Scholar]
- 160. Drewes G, Ebneth A, Preuss U, Mandelkow EM, Mandelkow E (1997) MARK, a novel family of protein kinases that phosphorylate microtubule‐associated proteins and trigger microtubule disruption. Cell 89:297–308. [DOI] [PubMed] [Google Scholar]
- 161. Reynolds CH, Utton MA, Gibb GM, Yates A, Anderton BH (1997) Stress‐activated protein kinase/c‐Jun N‐terminal kinase phosphorylates tau protein. J Neurochem 68:1736–1744. [DOI] [PubMed] [Google Scholar]
- 162. Liu F, Grundke‐Iqbal I, Iqbal K, Gong CX (2005) Contributions of protein phosphatases PP1, PP2A, PP2B and PP5 to the regulation of tau phosphorylation. Eur J Neurosci 22:1942–1950. [DOI] [PubMed] [Google Scholar]
- 163. Martin L, Latypova X, Wilson CM, Magnaudeix A, Perrin ML, Terro F (2013) Tau protein phosphatases in Alzheimer's disease: the leading role of PP2A. Ageing Res Rev 12:39–49. [DOI] [PubMed] [Google Scholar]
- 164. Haase C, Stieler JT, Arendt T, Holzer M (2004) Pseudophosphorylation of tau protein alters its ability for self‐aggregation. J Neurochem 88:1509–1520. [DOI] [PubMed] [Google Scholar]
- 165. Brunden KR, Trojanowski JQ, Lee VM (2009) Advances in tau‐focused drug discovery for Alzheimer's disease and related tauopathies. Nat Rev Drug Discov 8:783–793. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 166. Khlistunova I, Biernat J, Wang Y, Pickhardt M, von Bergen M, Gazova Z, Mandelkow E, Mandelkow EM (2006) Inducible expression of Tau repeat domain in cell models of tauopathy: aggregation is toxic to cells but can be reversed by inhibitor drugs. J Biol Chem 281:1205–1214. [DOI] [PubMed] [Google Scholar]
- 167. Ando K, Maruko‐Otake A, Ohtake Y, Hayashishita M, Sekiya M, Iijima KM (2016) Stabilization of microtubule‐unbound Tau via Tau phosphorylation at Ser262/356 by Par‐1/MARK contributes to augmentation of AD‐related phosphorylation and Aβ42‐induced Tau toxicity. PLoS Genet 12:e1005917. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 168. Schneider A, Biernat J, von Bergen M, Mandelkow E, Mandelkow EM (1999) Phosphorylation that detaches tau protein from microtubules (Ser262, Ser214) also protects it against aggregation into Alzheimer paired helical filaments. Biochemistry 38:3549–3558. [DOI] [PubMed] [Google Scholar]
- 169. Kenessey A, Yen SH (1993) The extent of phosphorylation of fetal tau is comparable to that of PHF‐tau from Alzheimer paired helical filaments. Brain Res 629:40–46. [DOI] [PubMed] [Google Scholar]
- 170. Härtig W, Stieler J, Boerema AS, Wolf J, Schmidt U, Weissfuss J, Bullmann T, Strijkstra AM, Arendt T (2007) Hibernation model of tau phosphorylation in hamsters: selective vulnerability of cholinergic basal forebrain neurons ‐ implications for Alzheimer's disease. Eur J Neurosci 25:69–80. [DOI] [PubMed] [Google Scholar]
- 171. Lee S, Shea TB (2012) Caspase‐mediated truncation of Tau potentiates aggregation. Int J Alzheimers Dis 2012:731063. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 172. Park SY, Ferreira A (2005) The generation of a 17 kDa neurotoxic fragment: an alternative mechanism by which tau mediates beta‐amyloid‐induced neurodegeneration. J Neurosci 25:5365–5375. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 173. Corsetti V, Amadoro G, Gentile A, Capsoni S, Ciotti MT, Cencioni MT, Atlante A, Canu N, Rohn TT, Cattaneo A, Calissano P (2008) Identification of a caspase‐derived N‐terminal tau fragment in cellular and animal Alzheimer's disease models. Mol Cell Neurosci 38:381–392. [DOI] [PubMed] [Google Scholar]
- 174. Johnson GV, Seubert P, Cox TM, Motter R, Brown JP, Galasko D (1997) The tau protein in human cerebrospinal fluid in Alzheimer's disease consists of proteolytically derived fragments. J Neurochem 68:430–433. [DOI] [PubMed] [Google Scholar]
- 175. Matsumoto SE, Motoi Y, Ishiguro K, Tabira T, Kametani F, Hasegawa M, Hattori N (2015) The twenty‐four KDa C‐terminal tau fragment increases with aging in tauopathy mice: implications of prion‐like properties. Hum Mol Genet 24:6403–6416. [DOI] [PubMed] [Google Scholar]
- 176. Nübling G, Levin J, Bader B, Israel L, Bötzel K, Lorenzl S, Giese A (2012) Limited cleavage of tau with matrix‐metalloproteinase MMP‐9, but not MMP‐3, enhances tau oligomer formation. Exp Neurol 237:470–476. [DOI] [PubMed] [Google Scholar]
- 177. Skrabana R, Sevcik J, Novak M (2006) Intrinsically disordered proteins in the neurodegenerative processes: formation of tau protein paired helical filaments and their analysis. Cell Mol Neurobiol 26:1085–1097. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 178. Wang YP, Biernat J, Pickhardt M, Mandelkow E, Mandelkow EM (2007) Stepwise proteolysis liberates tau fragments that nucleate the Alzheimer‐like aggregation of full‐length tau in a neuronal cell model. Proc Natl Acad Sci USA 104:10252–10257. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 179. Novak M, Kabat J, Wischik CM (1993) Molecular characterization of the minimal protease resistant tau unit of the Alzheimer's disease paired helical filament. EMBO J 12:365–370. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 180. Basurto‐Islas G, Luna‐Muñoz J, Guillozet‐Bongaarts AL, Binder LI, Mena R, García‐Sierra F (2008) Accumulation of aspartic acid421‐ and glutamic acid391‐cleaved tau in neurofibrillary tangles correlates with progression in Alzheimer disease. J Neuropathol Exp Neurol 67:470–483. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 181. McMillan PJ, Kraemer BC, Robinson L, Leverenz JB, Raskind M, Schellenberg G (2011) Truncation of tau at E391 promotes early pathologic changes in transgenic mice. J Neuropathol Exp Neurol 70:1006–1019. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 182. Liu F, Zaidi T, Iqbal K, Grundke‐Iqbal I, Gong CX (2002) Aberrant glycosylation modulates phosphorylation of tau by protein kinase A and dephosphorylation of tau by protein phosphatase 2A and 5. Neuroscience 115:829–837. [DOI] [PubMed] [Google Scholar]
- 183. Wang JZ, Grundke‐Iqbal I, Iqbal K (1996) Glycosylation of microtubule‐associated protein tau: an abnormal posttranslational modification in Alzheimer's disease. Nat Med 2:871–875. [DOI] [PubMed] [Google Scholar]
- 184. Lefebvre T, Ferreira S, Dupont‐Wallois L, Bussière T, Dupire MJ, Delacourte A, Michalski JC, Caillet‐Boudin ML (2003) Evidence of a balance between phosphorylation and O‐GlcNAc glycosylation of Tau proteins–a role in nuclear localization. Biochim Biophys Acta 1619:167–176. [DOI] [PubMed] [Google Scholar]
- 185. Schedin‐Weiss S, Winblad B, Tjernberg LO (2014) The role of protein glycosylation in Alzheimer disease. FEBS J 281:46–62. [DOI] [PubMed] [Google Scholar]
- 186. Robertson LA, Moya KL, Breen KC (2004) The potential role of tau protein O‐glycosylation in Alzheimer's disease. J Alzheimers Dis 6:489–495. [DOI] [PubMed] [Google Scholar]
- 187. Liu F, Iqbal K, Grundke‐Iqbal I, Hart GW, Gong CX (2004) O‐GlcNAcylation regulates phosphorylation of tau: a mechanism involved in Alzheimer's disease. Proc Natl Acad Sci USA 101:10804–10809. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 188. Irwin DJ, Cohen TJ, Grossman M, Arnold SE, Xie SX, Lee VM, Trojanowski JQ (2012) Acetylated tau, a novel pathological signature in Alzheimer's disease and other tauopathies. Brain 135:807–818. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 189. Trzeciakiewicz H, Tseng JH, Wander CM, Madden V, Tripathy A, Yuan CX, Cohen TJ (2017) A dual pathogenic mechanism links Tau acetylation to sporadic tauopathy. Sci Rep 7:44102. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 190. Liou YC, Sun A, Ryo A, Zhou XZ, Yu ZX, Huang HK, Uchida T, Bronson R, Bing G, Li X, Hunter T, Lu KP (2003) Role of the prolyl isomerase Pin1 in protecting against age‐dependent neurodegeneration. Nature 424:556–561. [DOI] [PubMed] [Google Scholar]
- 191. Kondo A, Shahpasand K, Mannix R, Qiu J, Moncaster J, Chen CH, Yao Y, Lin YM, Driver JA, Sun Y, Wei S, Luo ML, Albayram O, Huang P, Rotenberg A, Ryo A, Goldstein LE, Pascual‐Leone A, McKee AC, Meehan W, Zhou XZ, Lu KP (2015) Antibody against early driver of neurodegeneration cis P‐tau blocks brain injury and tauopathy. Nature 523:431–436. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 192. Tai HC, Serrano‐Pozo A, Hashimoto T, Frosch MP, Spires‐Jones TL, Hyman BT (2012) The synaptic accumulation of hyperphosphorylated tau oligomers in Alzheimer disease is associated with dysfunction of the ubiquitin‐proteasome system. Am J Pathol 181:1426–1435. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 193. Halverson RA, Lewis J, Frausto S, Hutton M, Muma NA (2005) Tau protein is cross‐linked by transglutaminase in P301L tau transgenic mice. J Neurosci 25:1226–1233. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 194. Norlund MA, Lee JM, Zainelli GM, Muma NA (1999) Elevated transglutaminase‐induced bonds in PHF tau in Alzheimer's disease. Brain Res 851:154–163. [DOI] [PubMed] [Google Scholar]
- 195. Zemaitaitis MO, Lee JM, Troncoso JC, Muma NA (2000) Transglutaminase‐induced cross‐linking of tau proteins in progressive supranuclear palsy. J Neuropathol Exp Neurol 59:983–989. [DOI] [PubMed] [Google Scholar]
- 196. Tucholski J, Kuret J, Johnson GV (1999) Tau is modified by tissue transglutaminase in situ: possible functional and metabolic effects of polyamination. J Neurochem 73:1871–1880. [PubMed] [Google Scholar]
- 197. Necula M, Kuret J (2004) Pseudophosphorylation and glycation of tau protein enhance but do not trigger fibrillization in vitro. J Biol Chem 279:49694–49703. [DOI] [PubMed] [Google Scholar]
- 198. Ledesma MD, Bonay P, Colaço C, Avila J (1994) Analysis of microtubule‐associated protein tau glycation in paired helical filaments. J Biol Chem 269:21614–21619. [PubMed] [Google Scholar]
- 199. Zhang YJ, Xu YF, Chen XQ, Wang XC, Wang JZ (2005) Nitration and oligomerization of tau induced by peroxynitrite inhibit its microtubule‐binding activity. FEBS Lett 579:2421–2427. [DOI] [PubMed] [Google Scholar]
- 200. Maeda S, Sahara N, Saito Y, Murayama M, Yoshiike Y, Kim H, Miyasaka T, Murayama S, Ikai A, Takashima A (2007) Granular tau oligomers as intermediates of tau filaments. Biochemistry 46:3856–3861. [DOI] [PubMed] [Google Scholar]
- 201. Kayed R, Sokolov Y, Edmonds B, McIntire TM, Milton SC, Hall JE, Glabe CG (2004) Permeabilization of lipid bilayers is a common conformation‐dependent activity of soluble amyloid oligomers in protein misfolding diseases. J Biol Chem 279:46363–46366. [DOI] [PubMed] [Google Scholar]
- 202. Quist A, Doudevski I, Lin H, Azimova R, Ng D, Frangione B, Kagan B, Ghiso J, Lal R (2005) Amyloid ion channels: a common structural link for protein‐misfolding disease. Proc Natl Acad Sci USA 102:10427–10432. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 203. Fang YS, Tsai KJ, Chang YJ, Kao P, Woods R, Kuo PH, Wu CC, Liao JY, Chou SC, Lin V, Jin LW, Yuan HS, Cheng IH, Tu PH, Chen YR (2014) Full‐length TDP‐43 forms toxic amyloid oligomers that are present in frontotemporal lobar dementia‐TDP patients. Nat Commun 5:4824. [DOI] [PubMed] [Google Scholar]
- 204. Hawkins BE, Krishnamurthy S, Castillo‐Carranza DL, Sengupta U, Prough DS, Jackson GR, DeWitt DS, Kayed R (2013) 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 288:17042–17050. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 205. Violet M, Chauderlier A, Delattre L, Tardivel M, Chouala MS, Sultan A, Marciniak E, Humez S, Binder L, Kayed R, Lefebvre B, Bonnefoy E, Buée L, Galas MC (2015) Prefibrillar Tau oligomers alter the nucleic acid protective function of Tau in hippocampal neurons in vivo. Neurobiol Dis 82:540–551. [DOI] [PubMed] [Google Scholar]
- 206. Maeda S, Sahara N, Saito Y, Murayama S, Ikai A, Takashima A (2006) Increased levels of granular tau oligomers: an early sign of brain aging and Alzheimer's disease. Neurosci Res 54:197–201. [DOI] [PubMed] [Google Scholar]
- 207. Lasagna‐Reeves CA, Castillo‐Carranza DL, Sengupta U, Clos AL, Jackson GR, Kayed R (2011) Tau oligomers impair memory and induce synaptic and mitochondrial dysfunction in wild‐type mice. Mol Neurodegener 6:39. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 208. Lasagna‐Reeves CA, Castillo‐Carranza DL, Sengupta U, Guerrero‐Munoz MJ, Kiritoshi T, Neugebauer V, Jackson GR, Kayed R (2012) Alzheimer brain‐derived tau oligomers propagate pathology from endogenous tau. Sci Rep 2:700. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 209. Fá M, Puzzo D, Piacentini R, Staniszewski A, Zhang H, Baltrons MA, Li Puma DD, Chatterjee I, Li J, Saeed F, Berman HL, Ripoli C, Gulisano W, Gonzalez J, Tian H, Costa JA, Lopez P, Davidowitz E, Yu WH, Haroutunian V, Brown LM, Palmeri A, Sigurdsson EM, Duff KE, Teich AF, Honig LS, Sierks M, Moe JG, D'Adamio L, Grassi C, Kanaan NM, Fraser PE, Arancio O (2016) Extracellular tau oligomers produce an immediate impairment of LTP and memory. Sci Rep 6:19393. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 210. Cowan CM, Quraishe S, Hands S, Sealey M, Mahajan S, Allan DW, Mudher A (2015) Rescue from tau‐induced neuronal dysfunction produces insoluble tau oligomers. Sci Rep 5:17191. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 211. Do TD, Economou NJ, Chamas A, Buratto SK, Shea JE, Bowers MT (2014) Interactions between amyloid‐β and tau fragments promote aberrant aggregates: implications for amyloid toxicity. J Phys Chem B 118:11220–11230. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 212. Bieschke J, Herbst M, Wiglenda T, Friedrich RP, Boeddrich A, Schiele F, Kleckers D, Lopez del Amo JM, Grüning BA, Wang Q, Schmidt MR, Lurz R, Anwyl R, Schnoegl S, Fändrich M, Frank RF, Reif B, Günther S, Walsh DM, Wanker EE (2011) Small‐molecule conversion of toxic oligomers to nontoxic β‐sheet–rich amyloid fibrils. Nat Chem Biol 8:93–101. [DOI] [PubMed] [Google Scholar]
- 213. Campioni S, Mannini B, Zampagni M, Pensalfini A, Parrini C, Evangelisti E, Relini A, Stefani M, Dobson CM, Cecchi C, Chiti F (2010) A causative link between the structure of aberrant protein oligomers and their toxicity. Nat Chem Biol 6:140–147. [DOI] [PubMed] [Google Scholar]
- 214. Kfoury N, Holmes BB, Jiang H, Holtzman DM, Diamond MI (2012) Trans‐cellular propagation of Tau aggregation by fibrillar species. J Biol Chem 287:19440–19451. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 215. Iba M, Guo JL, McBride JD, Zhang B, Trojanowski JQ, Lee VM (2013) Synthetic tau fibrils mediate transmission of neurofibrillary tangles in a transgenic mouse model of Alzheimer's‐like tauopathy. J Neurosci 33:1024–1037. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 216. Iqbal K, Liu F, Gong CX, Grundke‐Iqbal I (2010) Tau in Alzheimer disease and related tauopathies. Curr Alzheimer Res 7:656–664. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 217. Crowther RA (1991) Straight and paired helical filaments in Alzheimer disease have a common structural unit. Proc Natl Acad Sci USA 88:2288–2292. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 218. Spillantini MG, Goedert M (1998) Tau protein pathology in neurodegenerative diseases. Trends Neurosci 21:428–433. [DOI] [PubMed] [Google Scholar]
- 219. Barghorn S, Davies P, Mandelkow E (2004) Tau paired helical filaments from Alzheimer's disease brain and assembled in vitro are based on beta‐structure in the core domain. Biochemistry 43:1694–1703. [DOI] [PubMed] [Google Scholar]
- 220. King ME, Ahuja V, Binder LI, Kuret J (1999) Ligand‐dependent tau filament formation: implications for Alzheimer's disease progression. Biochemistry 38:14851–14859. [DOI] [PubMed] [Google Scholar]
- 221. Wischik CM, Novak M, Thøgersen HC, Edwards PC, Runswick MJ, Jakes R, Walker JE, Milstein C, Roth M, Klug A (1988) Isolation of a fragment of tau derived from the core of the paired helical filament of Alzheimer disease. Proc Natl Acad Sci USA 85:4506–4510. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 222. Ksiezak‐Reding H, Wall JS (2005) Characterization of paired helical filaments by scanning transmission electron microscopy. Microsc Res Tech 67:126–140. [DOI] [PubMed] [Google Scholar]
- 223. Wegmann S, Medalsy ID, Mandelkow E, Müller DJ (2013) The fuzzy coat of pathological human Tau fibrils is a two‐layered polyelectrolyte brush. Proc Natl Acad Sci USA 110:E313–E321. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 224. Goedert M (2015) Neurodegeneration. Alzheimer's and Parkinson's diseases: the prion concept in relation to assembled Aβ, tau, and α‐synuclein. Science 349:1255555. [DOI] [PubMed] [Google Scholar]
- 225. Kouri N, Whitwell JL, Josephs KA, Rademakers R, Dickson DW (2011) Corticobasal degeneration: a pathologically distinct 4R tauopathy. Nat Rev Neurol 7:263–272. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 226. Bronner IF, ter Meulen BC, Azmani A, Severijnen LA, Willemsen R, Kamphorst W, Ravid R, Heutink P, van Swieten JC (2005) Hereditary Pick's disease with the G272V tau mutation shows predominant three‐repeat tau pathology. Brain 128:2645–2653. [DOI] [PubMed] [Google Scholar]
- 227. Murayama S, Mori H, Ihara Y, Tomonaga M (1990) Immunocytochemical and ultrastructural studies of Pick's disease. Ann Neurol 27:394–405. [DOI] [PubMed] [Google Scholar]
- 228. Wegmann S, Jung YJ, Chinnathambi S, Mandelkow EM, Mandelkow E, Muller DJ (2010) Human tau isoforms assemble into ribbon‐like fibrils that display polymorphic structure and stability. J Biol Chem 285:27302–27313. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 229. Daebel V, Chinnathambi S, Biernat J, Schwalbe M, Habenstein B, Loquet A, Akoury E, Tepper K, Müller H, Baldus M, Griesinger C, Zweckstetter M, Mandelkow E, Vijayan V, Lange A (2012) β‐sheet core of tau paired helical filaments revealed by solid‐state NMR. J Am Chem Soc 134:13982–13989. [DOI] [PubMed] [Google Scholar]
- 230. Berriman J, Serpell LC, Oberg KA, Fink AL, Goedert M, Crowther RA (2003) Tau filaments from human brain and from in vitro assembly of recombinant protein show cross‐beta structure. Proc Natl Acad Sci USA 100:9034–9038. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 231. Sadqi M, Hernández F, Pan U, Pérez M, Schaeberle MD, Avila J, Muñoz V (2002) Alpha‐helix structure in Alzheimer's disease aggregates of tau‐protein. Biochemistry 41:7150–7155. [DOI] [PubMed] [Google Scholar]
- 232. Goux WJ, Kopplin L, Nguyen AD, Leak K, Rutkofsky M, Shanmuganandam VD, Sharma D, Inouye H, Kirschner DA (2004) The formation of straight and twisted filaments from short tau peptides. J Biol Chem 279:26868–26875. [DOI] [PubMed] [Google Scholar]
- 233. Xu L, Zheng J, Margittai M, Nussinov R, Ma B (2016) How does hyperphopsphorylation promote tau aggregation and modulate filament structure and stability?. ACS Chem Neurosci 7:565–575. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 234. Morales R, Moreno‐Gonzalez I, Soto C (2013) Cross‐seeding of misfolded proteins: implications for etiology and pathogenesis of protein misfolding diseases. PLoS Pathog 9:e1003537. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 235. Soto C (2012) Transmissible proteins: expanding the prion heresy. Cell 149:968–977. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 236. Morales R, Estrada LD, Diaz‐Espinoza R, Morales‐Scheihing D, Jara MC, Castilla J, Soto C (2010) Molecular cross talk between misfolded proteins in animal models of Alzheimer's and prion diseases. J Neurosci 30:4528–4535. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 237. Vasconcelos B, Stancu IC, Buist A, Bird M, Wang P, Vanoosthuyse A, Van Kolen K, Verheyen A, Kienlen‐Campard P, Octave JN, Baatsen P, Moechars D, Dewachter I (2016) Heterotypic seeding of Tau fibrillization by pre‐aggregated Abeta provides potent seeds for prion‐like seeding and propagation of Tau‐pathology in vivo. Acta Neuropathol 131:549–569. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 238. Ferrari A, Hoerndli F, Baechi T, Nitsch RM, Gotz J (2003) beta‐amyloid induces PHF‐like tau filaments in tissue culture. J Biol Chem 278:40162–40168. [DOI] [PubMed] [Google Scholar]
- 239. Guo JP, Arai T, Miklossy J, McGeer PL (2006) Abeta and tau form soluble complexes that may promote self aggregation of both into the insoluble forms observed in Alzheimer's disease. Proc Natl Acad Sci USA 103:1953–1958. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 240. Miller Y, Ma B, Nussinov R (2011) Synergistic interactions between repeats in tau protein and Aβ amyloids may be responsible for accelerated aggregation via polymorphic states. Biochemistry 50:5172–5181. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 241. Lewis J, Dickson DW, Lin WL, Chisholm L, Corral A, Jones G, Yen SH, Sahara N, Skipper L, Yager D, Eckman C, Hardy J, Hutton M, McGowan E (2001) Enhanced neurofibrillary degeneration in transgenic mice expressing mutant tau and APP. Science 293:1487–1491. [DOI] [PubMed] [Google Scholar]
- 242. Götz J, Chen F, van Dorpe J, Nitsch RM (2001) Formation of neurofibrillary tangles in P301L tau transgenic mice induced by abeta 42 fibrils. Science 293:1491–1495. [DOI] [PubMed] [Google Scholar]
- 243. Qi R, Luo Y, Wei G, Nussinov R, Ma B (2015) Aβ “stretching‐and‐packing” cross‐seeding mechanism can trigger tau protein aggregation. J Phys Chem Lett 6:3276–3282. [Google Scholar]
- 244. Jensen PH, Hager H, Nielsen MS, Højrup P, Gliemann J, Jakes R (1999) Alpha‐synuclein binds to tau and stimulates the protein kinase A‐catalyzed tau phosphorylation of serine residues 262 and 356. J Biol Chem 274:25481–25489. [DOI] [PubMed] [Google Scholar]
- 245. Waxman EA, Giasson BI (2011) Induction of intracellular tau aggregation is promoted by α‐synuclein seeds and provides novel insights into the hyperphosphorylation of tau. J Neurosci 31:7604–7618. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 246. Duda JE, Lee VM, Trojanowski JQ (2000) Neuropathology of synuclein aggregates: new insights into mechanisms of neurodegenerative diseases. J Neurosci Res 61:121–127. [DOI] [PubMed] [Google Scholar]
- 247. Bachhuber T, Katzmarski N, McCarter JF, Loreth D, Tahirovic S, Kamp F, Abou‐Ajram C, Nuscher B, Serrano‐Pozo A, Müller A, Prinz M, Steiner H, Hyman BT, Haass C, Meyer‐Luehmann M (2015) Inhibition of amyloid‐β plaque formation by α‐synuclein. Nat Med 21:802–807. [DOI] [PubMed] [Google Scholar]
- 248. Fernández‐Nogales M, Cabrera JR, Santos‐Galindo M, Hoozemans JJM, Ferrer I, Rozemuller AJ, Hernández F, Avila J, Lucas JJ (2014) Huntington's disease is a four‐repeat tauopathy with tau nuclear rods. Nat Med 20:881–885. [DOI] [PubMed] [Google Scholar]
- 249. Vuono R, Winder‐Rhodes S, de Silva R, Cisbani G, Drouin‐Ouellet J, Spillantini MG, Cicchetti F, Barker RA (2015) The role of tau in the pathological process and clinical expression of Huntington's disease. Brain 138:1907–1918. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 250. Blum D, Herrera F, Francelle L, Mendes T, Basquin M, Obriot H, Demeyer D, Sergeant N, Gerhardt E, Brouillet E, Buée L, Outeiro TF (2015) Mutant huntingtin alters Tau phosphorylation and subcellular distribution. Hum Mol Genet 24:76–85. [DOI] [PubMed] [Google Scholar]
- 251. Bautista MJ, Gutiérrez J, Salguero FJ, Fernández de Marco MM, Romero‐Trevejo JL, Gómez‐Villamandos JC (2006) BSE infection in bovine PrP transgenic mice leads to hyperphosphorylation of tau‐protein. Vet Microbiol 115:293–301. [DOI] [PubMed] [Google Scholar]
- 252. Ghetti B, Piccardo P, Spillantini MG, Ichimiya Y, Porro M, Perini F, Kitamoto T, Tateishi J, Seiler C, Frangione B, Bugiani O, Giaccone G, Prelli F, Goedert M, Dlouhy SR, Tagliavini F (1996) Vascular variant of prion protein cerebral amyloidosis with tau‐positive neurofibrillary tangles: the phenotype of the stop codon 145 mutation in PRNP. Proc Natl Acad Sci USA 93:744–748. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 253. Goodall CA, Head MW, Everington D, Ironside JW, Knight RS, Green AJ (2006) Raised CSF phospho‐tau concentrations in variant Creutzfeldt‐Jakob disease: diagnostic and pathological implications. J Neurol Neurosurg Psychiatry 77:89–91. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 254. Han J, Zhang J, Yao H, Wang X, Li F, Chen L, Gao C, Gao J, Nie K, Zhou W, Dong X (2006) Study on interaction between microtubule associated protein tau and prion protein. Sci China C Life Sci 49:473–479. [DOI] [PubMed] [Google Scholar]
- 255. Wang XF, Dong CF, Zhang J, Wan YZ, Li F, Huang YX, Han L, Shan B, Gao C, Han J, Dong XP (2008) Human tau protein forms complex with PrP and some GSS‐ and fCJD‐related PrP mutants possess stronger binding activities with tau in vitro. Mol Cell Biochem 310:49–55. [DOI] [PubMed] [Google Scholar]
- 256. Clavaguera F, Akatsu H, Fraser G, Crowther RA, Frank S, Hench J, Probst A, Winkler DT, Reichwald J, Staufenbiel M, Ghetti B, Goedert M, Tolnay M (2013) Brain homogenates from human tauopathies induce tau inclusions in mouse brain. Proc Natl Acad Sci USA 110:9535–9540. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 257. Surmacz‐Chwedoruk W, Nieznańska H, Wójcik S, Dzwolak W (2012) Cross‐seeding of fibrils from two types of insulin induces new amyloid strains. Biochemistry 51:9460–9469. [DOI] [PubMed] [Google Scholar]
- 258. Petkova AT, Leapman RD, Guo Z, Yau WM, Mattson MP, Tycko R (2005) Self‐propagating, molecular‐level polymorphism in Alzheimer's beta‐amyloid fibrils. Science 307:262–265. [DOI] [PubMed] [Google Scholar]
- 259. Shorter J, Lindquist S (2005) Prions as adaptive conduits of memory and inheritance. Nat Rev Genet 6:435–450. [DOI] [PubMed] [Google Scholar]
- 260. Chien P, Weissman JS, DePace AH (2004) Emerging principles of conformation‐based prion inheritance. Annu Rev Biochem 73:617–656. [DOI] [PubMed] [Google Scholar]
- 261. Jones EM, Surewicz WK (2005) Fibril conformation as the basis of species‐ and strain‐dependent seeding specificity of mammalian prion amyloids. Cell 121:63–72. [DOI] [PubMed] [Google Scholar]
- 262. Yamaguchi K, Takahashi S, Kawai T, Naiki H, Goto Y (2005) Seeding‐dependent propagation and maturation of amyloid fibril conformation. J Mol Biol 352:952–960. [DOI] [PubMed] [Google Scholar]
- 263. Morozova OA, March ZM, Robinson AS, Colby DW (2013) Conformational features of tau fibrils from Alzheimer's disease brain are faithfully propagated by unmodified recombinant protein. Biochemistry 52:6960–6967. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 264. Dinkel PD, Siddiqua A, Huynh H, Shah M, Margittai M (2011) Variations in filament conformation dictate seeding barrier between three‐ and four‐repeat tau. Biochemistry 50:4330–4336. [DOI] [PubMed] [Google Scholar]
- 265. Yu X, Luo Y, Dinkel P, Zheng J, Wei G, Margittai M, Nussinov R, Ma B (2012) Cross‐seeding and conformational selection between three‐ and four‐repeat human Tau proteins. J Biol Chem 287:14950–14959. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 266. Meyer V, Holden MR, Weismiller HA, Eaton GR, Eaton SS, Margittai M (2016) Fracture and growth are competing forces determining the fate of conformers in tau fibril populations. J Biol Chem 291:12271–12281. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 267. Hall GF, Patuto BA (2012) Is tau ready for admission to the prion club? Prion 6:223–233. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 268. Baskakov IV (2009) Switching in amyloid structure within individual fibrils: implication for strain adaptation, species barrier and strain classification. FEBS Lett 583:2618–2622. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 269. Collinge J (2016) Mammalian prions and their wider relevance in neurodegenerative diseases. Nature 539:217–226. [DOI] [PubMed] [Google Scholar]
- 270. Li J, Browning S, Mahal SP, Oelschlegel AM, Weissmann C (2010) Darwinian evolution of prions in cell culture. Science 327:869–872. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 271. Makarava N, Baskakov IV (2013) The evolution of transmissible prions: the role of deformed templating. PLoS Pathog 9:e1003759. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 272. Makarava N, Ostapchenko VG, Savtchenko R, Baskakov IV (2009) Conformational switching within individual amyloid fibrils. J Biol Chem 284:14386–14395. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 273. Aoyagi H, Hasegawa M, Tamaoka A (2007) Fibrillogenic nuclei composed of P301L mutant tau induce elongation of P301L tau but not wild‐type tau. J Biol Chem 282:20309–20318. [DOI] [PubMed] [Google Scholar]
- 274. von Bergen M, Barghorn S, Biernat J, Mandelkow EM, Mandelkow E (2005) Tau aggregation is driven by a transition from random coil to beta sheet structure. Biochim Biophys Acta 1739:158–166. [DOI] [PubMed] [Google Scholar]
- 275. Prusiner SB (2012) Cell biology. A unifying role for prions in neurodegenerative diseases. Science 336:1511–1513. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 276. Clavaguera F, Bolmont T, Crowther RA, Abramowski D, Frank S, Probst A, Fraser G, Stalder AK, Beibel M, Staufenbiel M, Jucker M, Goedert M, Tolnay M (2009) Transmission and spreading of tauopathy in transgenic mouse brain. Nat Cell Biol 11:909–913. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 277. Peeraer E, Bottelbergs A, Van Kolen K, Stancu IC, Vasconcelos B, Mahieu M, Duytschaever H, Ver Donck L, Torremans A, Sluydts E, Van Acker N, Kemp JA, Mercken M, Brunden KR, Trojanowski JQ, Dewachter I, Lee VM, Moechars D (2015) Intracerebral injection of preformed synthetic tau fibrils initiates widespread tauopathy and neuronal loss in the brains of tau transgenic mice. Neurobiol Dis 73:83–95. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 278. Wu JW, Herman M, Liu L, Simoes S, Acker CM, Figueroa H, Steinberg JI, Margittai M, Kayed R, Zurzolo C, Di Paolo G, Duff KE (2013) Small misfolded tau species are internalized via bulk endocytosis and anterogradely and retrogradely transported in neurons. J Biol Chem 288:1856–1870. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 279. Frost B, Jacks RL, Diamond MI (2009) Propagation of Tau misfolding from the outside to the inside of a cell. J Biol Chem 284:12845–12852. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 280. Guo JL, Lee VM (2014) Cell‐to‐cell transmission of pathogenic proteins in neurodegenerative diseases. Nat Med 20:130–138. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 281. Wegmann S, Maury EA, Kirk MJ, Saqran L, Roe A, DeVos SL, Nicholls S, Fan Z, Takeda S, Cagsal‐Getkin O, William CM, Spires‐Jones TL, Pitstick R, Carlson GA, Pooler AM, Hyman BT (2015) Removing endogenous tau does not prevent tau propagation yet reduces its neurotoxicity. EMBO J 34:3028–3041. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 282. Brandner S, Isenmann S, Raeber A, Fischer M, Sailer A, Kobayashi Y, Marino S, Weissmann C, Aguzzi A (1996) Normal host prion protein necessary for scrapie‐induced neurotoxicity. Nature 379:339–343. [DOI] [PubMed] [Google Scholar]
- 283. Aguzzi A, Lakkaraju AK (2016) Cell biology of prions and prionoids: a status report. Trends Cell Biol 26:40–51. [DOI] [PubMed] [Google Scholar]
- 284. Aguzzi A, Rajendran L (2009) The transcellular spread of cytosolic amyloids, prions, and prionoids. Neuron 64:783–790. [DOI] [PubMed] [Google Scholar]
- 285. Beekes M, Thomzig A, Schulz‐Schaeffer WJ, Burger R (2014) Is there a risk of prion‐like disease transmission by Alzheimer‐ or Parkinson‐associated protein particles?. Acta Neuropathol 128:463–476. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 286. Barron RM, King D, Jeffrey M, McGovern G, Agarwal S, Gill AC, Piccardo P (2016) PrP aggregation can be seeded by pre‐formed recombinant PrP amyloid fibrils without the replication of infectious prions. Acta Neuropathol 132:611–624. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 287. Boutajangout A, Ingadottir J, Davies P, Sigurdsson EM (2011) Passive immunization targeting pathological phospho‐tau protein in a mouse model reduces functional decline and clears tau aggregates from the brain. J Neurochem 118:658–667. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 288. Noble W, Planel E, Zehr C, Olm V, Meyerson J, Suleman F, Gaynor K, Wang L, LaFrancois J, Feinstein B, Burns M, Krishnamurthy P, Wen Y, Bhat R, Lewis J, Dickson D, Duff K (2005) Inhibition of glycogen synthase kinase‐3 by lithium correlates with reduced tauopathy and degeneration in vivo. Proc Natl Acad Sci USA 102:6990–6995. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 289. van Eersel J, Ke YD, Liu X, Delerue F, Kril JJ, Gotz J, Ittner LM (2010) Sodium selenate mitigates tau pathology, neurodegeneration, and functional deficits in Alzheimer's disease models. Proc Natl Acad Sci USA 107:13888–13893. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 290. Zhang B, Maiti A, Shively S, Lakhani F, McDonald‐Jones G, Bruce J, Lee EB, Xie SX, Joyce S, Li C, Toleikis PM, Lee VM, Trojanowski JQ (2005) Microtubule‐binding drugs offset tau sequestration by stabilizing microtubules and reversing fast axonal transport deficits in a tauopathy model. Proc Natl Acad Sci USA 102:227–231. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 291. Yanamandra K, Kfoury N, Jiang H, Mahan TE, Ma S, Maloney SE, Wozniak DF, Diamond MI, Holtzman DM (2013) Anti‐tau antibodies that block tau aggregate seeding in vitro markedly decrease pathology and improve cognition in vivo. Neuron 80:402–414. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 292. Kadavath H, Jaremko M, Jaremko Ł, Biernat J, Mandelkow E, Zweckstetter M (2015) Folding of the tau protein on microtubules. Angew Chem Int Ed Engl 54:10347–10351. [DOI] [PubMed] [Google Scholar]
- 293. DeLano W (2002) The PyMOL Molecular Graphics System. De‐Lano Scientific, San Carlos, CA, USA. http//www.pymol.org.
