Entry - *157140 - MICROTUBULE-ASSOCIATED PROTEIN TAU; MAPT - OMIM
* 157140

MICROTUBULE-ASSOCIATED PROTEIN TAU; MAPT


Alternative titles; symbols

MTBT1


HGNC Approved Gene Symbol: MAPT

Cytogenetic location: 17q21.31     Genomic coordinates (GRCh38): 17:45,894,554-46,028,334 (from NCBI)


Gene-Phenotype Relationships
Location Phenotype Phenotype
MIM number
Inheritance Phenotype
mapping key
17q21.31 {Parkinson disease, susceptibility to} 168600 AD, Mu 3
Dementia, frontotemporal, with or without parkinsonism 600274 AD 3
Pick disease 172700 AD 3
Supranuclear palsy, progressive 601104 AD 3
Supranuclear palsy, progressive atypical 260540 AR 3

TEXT

Cloning and Expression

The microtubule-associated proteins (MAPs) coassemble with tubulin (see 602529) into microtubules in vitro. Microtubule-associated protein tau appears to be enriched in axons. Neve et al. (1986) identified tau cDNA clones in a human fetal brain cDNA library. The clones recognized a 6-kb message that was expressed in human brain but not in other human tissues and exhibited a developmental shift in size.

By screening cDNA libraries prepared from the frontal cortex of an Alzheimer disease patient and from fetal human brain, Goedert et al. (1988) isolated the cDNA for a core protein of the paired helical filament of Alzheimer disease (AD; 104300). The partial amino acid sequence of this core protein was used to design synthetic oligonucleotide probes. The cDNA encodes a protein of 352 amino acids that contains a characteristic amino acid repeat in its carboxyl-terminal half. Because of extensive homology to the sequence of the mouse microtubule-associated protein tau, they stated that this protein must constitute the human equivalent of mouse tau. Tau protein mRNA was found in normal amounts in the frontal cortex from patients with Alzheimer disease.

Goedert et al. (1989) determined the sequences of 6 tau isoforms produced in adult human brain by alternative mRNA splicing. The proteins are composed of 352 to 441 amino acids. The isoforms differ from each other by the presence or absence of 29-amino acid or 58-amino acid inserts located in the N terminus and a 31-amino repeat located in the C terminus. Inclusion of the latter, which is encoded by exon 10 of the tau gene, gives rise to the 3 tau isoforms with 4 repeats each; the other 3 isoforms have 3 repeats each. Normal cerebral cortex contains similar levels of 3-repeat and 4-repeat tau isoforms. The repeats and some adjoining sequences constitute the microtubule-binding domains of tau (see also Hutton et al., 1998).

Holzer et al. (2004) determined that the longest form (441 amino acids) of the chimpanzee brain tau protein shares 100% amino acid identity with human tau. The nonconstitutively spliced exon 4A differed at 3 of 251 amino acid positions. Identities for gorilla and gibbon tau with human tau were 99.5% and 99.0%, respectively. Analysis of 8 polymorphic markers revealed that the nonhuman primates had a higher prevalence of the H2 haplotype, in contrast to humans, in whom the H1 haplotype is more common. Holzer et al. (2004) noted that chimpanzee brains show resistance to developing tau pathology and suggested that differences in intronic sequences of the tau gene may be responsible.


Gene Structure

Andreadis et al. (1992) found that the tau isoforms that predominate in human brain are encoded by 11 exons.

Conrad et al. (2002) identified an intronless gene, saitohin (STH; 607067), located in the intron between exons 9 and 10 of the tau gene.


Mapping

The MAPT gene was assigned to chromosome 17 by hybridization of a cDNA clone to flow-sorted and spot-blotted chromosomes (Neve et al., 1986) and to 17q21 by in situ hybridization (Donlon et al., 1987).

In the course of constructing a radiation hybrid map of the breast cancer (113705) region of 17q, Abel et al. (1993) concluded that the tau protein gene (symbolized MTBT1 by them) is proximal to GP3A (173470), located at 17q21.32, and distal to TOP2A (126430) and THRA1 (190120), located at 17q11.2. ERBB2 (164870), previously localized to 17q21-q22, was found to be proximal to all of these. Poorkaj et al. (2001) identified and sequenced a human PAC and a mouse BAC containing the entire MAPT and Mtapt genes, respectively. They found that the corticotropin-releasing hormone receptor gene (CRHR1; 122561) is the next gene 5-prime to MAPT. The gene located 3-prime to MAPT and encoded by the opposite DNA strand has a predicted sequence identical to the KIAA1267 cDNA (KANSL1; 612452) identified from adult human brain by Nagase et al. (1999).


Gene Function

Alonso et al. (1996) reported studies on the microtubule-associated protein tau in Alzheimer disease. They noted that in the brains of patients with Alzheimer disease the neuronal cytoskeleton is progressively disrupted and replaced by tangles of paired helical filaments (PHFs), and that PHFs are composed mainly of hyperphosphorylated forms of tau (called 'AD P-tau' by them). They demonstrated that in solution normal tau associated with the hyperphosphorylated AD P-tau to form large tangles of filaments. They also demonstrated that dephosphorylation with alkaline phosphatase abolished the ability of AD P-tau to aggregate in vitro.

Hypothesizing that restoring the function of phosphorylated tau might prevent or reverse PHF formation in Alzheimer disease, and with the knowledge that PIN1 (601052) specifically isomerizes phosphorylation of a serine or threonine that precedes proline and regulates the function of mitotic phosphoproteins, Lu et al. (1999) demonstrated that the WW domain of PIN1 binds to phosphorylated tau at threonine-231 (T231). The T231 residue is hyperphosphorylated in Alzheimer disease and is phosphorylated to a certain extent in the normal brain. Using a pull-down assay, Lu et al. (1999) demonstrated that PIN1 binds to hyperphosphorylated tau from the brains of people with Alzheimer disease but not to tau from age-matched healthy brains. By immunoblotting, Lu et al. (1999) detected endogenous PIN1 in the PHFs of diseased brains, and using immunohistochemistry, they found that recombinant PIN1 binds to pathologic tau. Using immunohistochemistry, Lu et al. (1999) localized PIN1 to the nucleus in healthy brains. In the brains of people with Alzheimer disease, PIN1 staining was associated with pathologic tau in neuronal cells. Lu et al. (1999) also demonstrated that phosphorylated tau could neither bind microtubules nor promote microtubule assembly. However, PIN1 was able to restore the ability of phosphorylated tau to bind microtubules and promoted microtubule assembly in vitro. The level of soluble PIN1 in the brains of Alzheimer patients was greatly reduced compared to that in age-matched control brains. The authors concluded with the hypothesis that since depletion of PIN1 induces mitotic arrest and apoptotic cell death, sequestration of PIN1 into PHFs may contribute to neuronal death.

Schneider et al. (1999) presented evidence that challenged the common hypothesis that hyperphosphorylated tau promotes aggregation into PHFs. Using an in vitro system, they showed that treatment with several different kinases phosphorylated tau and caused detachment from microtubules, yet at the same time protected against the formation of PHFs. The authors suggested the findings are due most likely to the complex interaction of tau's ability to adopt many conformations and the presence of phosphorylation sites with differing affinities and locations.

Senile plaques and neurofibrillary tangles, the 2 hallmark lesions of Alzheimer disease, are the result of the pathologic deposition of proteins normally present throughout the brain. Senile plaques are extracellular deposits of fibrillar beta-amyloid peptide; neurofibrillary tangles represent intracellular bundles of self-assembled hyperphosphorylated tau proteins. These 2 lesions are often present in the same brain areas, and Rapoport et al. (2002) presented work that suggested a mechanistic link between them. They analyzed whether tau plays a key role in fibrillar beta-amyloid-induced neurite degeneration in central nervous system neurons. Cultured hippocampal neurons obtained from wildtype, tau knockout, and human tau transgenic mice were treated with fibrillar amyloid-beta. Morphologic analysis indicated that neurons expressing either mouse or human tau proteins degenerated in the presence of amyloid-beta. On the other hand, tau-depleted neurons showed no signs of degeneration in the presence of amyloid-beta.

From detailed analysis of pathologic load and spatiotemporal distribution of beta-amyloid deposits and tau pathology in sporadic Alzheimer disease, Delacourte et al. (2002) concluded that there is a synergetic effect of amyloid aggregation in the propagation of tau pathology.

The association of intronic mutations in the MAPT gene (e.g., 157140.0004) in frontotemporal dementia with parkinsonism (600274) highlights the involvement of aberrant pre-mRNA splicing in the pathogenesis of neurodegenerative disorders. To establish a model system for studying the role of pre-mRNA splicing in neurodegenerative diseases, Jiang et al. (2000) constructed a MAPT minigene that reproduced alternative splicing in both cultured cells and in vitro biochemical assays. They demonstrated that mutations in a nonconserved intronic region of the human MAPT gene led to increased splicing between exons 10 and 11. Systematic biochemical analyses indicated the importance of U1 snRNP (180740) and, to a lesser extent, U6 snRNP (180692) in differentially recognizing wildtype versus intron-mutant MAPT pre-mRNAs.

Goode et al. (2000) analyzed the structure and function of the 3-repeat (3R) and 4-repeat (4R) tau isoforms. Rather than having linear, sequentially arranged tubulin-binding domains, the isoforms have specific core microtubule-binding domains that lead to complex intramolecular folding interactions. Flanking regions were also found to contribute to the binding activity in the 3-repeat isoform, but less so in the 4-repeat isoform. Goode et al. (2000) suggested that the 2 isoforms form distinct structures that likely have different functional capabilities. Panda et al. (2003) noted that the abnormally high ratio of 4R to 3R tau in the MAPT gene might lead to neuronal cell death by altering normal tau functions in adult neurons. They tested whether 3R and 4R tau might differentially modulate the dynamic instability of microtubules in vitro using video microscopy. Although both isoforms promoted microtubule polymerization and decreased the tubulin critical subunit concentration to approximately similar extents, 4R tau stabilized microtubules significantly more strongly than 3R tau. Panda et al. (2003) suggested a 'dosage effect' or haploinsufficiency model in which both tau alleles must be active and properly regulated to produce appropriate amounts of each tau isoform to maintain microtubule dynamics within a tolerable window of activity.

Stamer et al. (2002) showed that elevated levels of tau inhibit intracellular transport in neurons, particularly the plus-end-directed transport by kinesin motors from the center of the cell body to the neuronal processes. This inhibition is significant because critical organelles, such as peroxisomes, mitochondria, and transport vesicles carrying supplies for the growth cone, are unable to penetrate the neurites, leading to stunted growth, increased susceptibility to oxidative stress, and likely pathologic aggregation of proteins such as amyloid precursor protein (APP; 104760). Stamer et al. (2002) concluded that the tau:tubulin ratio is normally low, and that increased levels of tau become detrimental to the cell.

Giasson et al. (2003) showed that alpha-synuclein (SNCA; 163890) induces fibrillization of tau, and that coincubation of alpha-synuclein and tau synergistically promotes fibrillization of both proteins in vitro. In vivo studies of mice with an alpha-synuclein mutation or a tau mutation showed filamentous inclusions of both proteins, which are abundant neuronal proteins that normally adopt an unfolded conformation but polymerize into amyloid fibrils in disease. The findings suggested an interaction between alpha-synuclein and tau that drives the formation of pathologic inclusions in human neurodegenerative diseases.

Rizzu et al. (2004) presented evidence suggesting that DJ1 (602533) colocalizes within a subset of pathologic tau inclusions in tauopathies, and that the solubility of DJ1 is altered in association with its aggregation within these inclusions.

Petrucelli et al. (2004) reported that CHIP (607207), a ubiquitin ligase that interacts directly with Hsp70/90 (140550/140571), induced ubiquitination and increased aggregation of tau. Tau lesions in human postmortem tissue were immunopositive for CHIP. Conversely, induction of Hsp70 through treatment with either geldanamycin or heat shock factor-1 (HSF1; 140580) led to a decrease in tau steady-state levels and a selective reduction in detergent insoluble tau. Furthermore, 30-month-old mice overexpressing inducible Hsp70 showed a significant reduction in tau levels. The authors concluded that the Hsp70/CHIP chaperone system may play an important role in the regulation of tau turnover and the selective elimination of abnormal tau species.

Liu et al. (2004) investigated the mechanisms leading to abnormal hyperphosphorylation of tau in pathologic states. They demonstrated that human brain tau was modified by O-GlcNAcylation, a type of protein O-glycosylation by which the monosaccharide beta-N-acetylglucosamine (GlcNAc) attaches to serine/threonine residues via an O-linked glycosidic bond. This glycosylation regulated tau phosphorylation in a site-specific manner both in vitro and in vivo. At most of the phosphorylation sites, the process negatively regulated tau phosphorylation. In an animal model of starved mice, low glucose uptake/metabolism that mimicked those observed in Alzheimer disease brain produced a decrease in O-GlcNAcylation and consequent hyperphosphorylation of tau at the majority of the phosphorylation sites. The O-GlcNAcylation level in Alzheimer disease brain extracts was decreased as compared to that in controls.

Using Western blotting, immunoprecipitation assays, and surface plasmon resonance analysis, Guo et al. (2006) showed that beta-amyloid-40 and -42 formed stable complexes with soluble tau and that prior phosphorylation of tau inhibited complex formation. Immunostaining of brain extracts from patients with AD and controls showed that phosphorylated tau and beta-amyloid were present within the same neuron. Guo et al. (2006) postulated that an initial step in AD pathogenesis may be the intracellular binding of soluble beta-amyloid to soluble nonphosphorylated tau.

In studies of rodent and human cells, Li et al. (2007) found that overexpression of hyperphosphorylated tau antagonized apoptosis of neuronal cells by stabilizing beta-catenin (CTNNB1; 116806). The findings explained why NFT-bearing neurons survive proapoptotic insults and instead die chronically of degeneration.

Roberson et al. (2007) found that reducing endogenous tau levels prevented behavioral deficits in transgenic mice expressing human APP, without altering their high A-beta levels. Tau reduction also protected both transgenic and nontransgenic mice against excitotoxicity. Roberson et al. (2007) concluded that thus, tau reduction can block A-beta- and excitotoxin-induced neuronal dysfunction and may represent an effective strategy for treating Alzheimer disease and related conditions.

To determine the effects of tau on dynein (600112) and kinesin (602809) motility, Dixit et al. (2008) conducted single-molecule studies of motor proteins moving along tau-decorated microtubules. Dynein tended to reverse direction, whereas kinesin tended to detach at patches of bound tau. Kinesin was inhibited at about a tenth of the tau concentration that inhibited dynein, and the microtubule-binding domain of tau was sufficient to inhibit motor activity. The differential modulation of dynein and kinesin motility suggested that microtubule-associated proteins (MAPs) can spatially regulate the balance of microtubule-dependent axonal transport.

De Calignon et al. (2010) used in vivo multiphoton imaging to observe tangles and activation of executioner caspases in living tau transgenic mice (Tg4510 strain) created by SantaCruz et al. (2005), and found that caspase activation occurs prior to tangle formation and precedes tangle formation by hours to days. New tangles form within a day. After a new tangle forms, the neuron remains alive and caspase activity seems to be suppressed. Similarly, introduction of wildtype 4-repeat tau (tau-4R) into wildtype animals triggered caspase activation, tau truncation, and tau aggregation. Adeno-associated virus-mediated expression of a construct mimicking caspase-cleaved tau into wildtype mice led to the appearance of intracellular aggregates, tangle-related conformational- and phospho-epitopes, and the recruitment of full-length endogenous tau to the aggregates. On the basis of these data, de Calignon et al. (2010) proposed a new model in which caspase activation cleaves tau to initiate tangle formation, then truncated tau recruits normal tau to misfold and form tangles. Because tangle-bearing neurons are long-lived, de Calignon et al. (2010) suggested that tangles are 'off pathway' to acute neuronal death. De Calignon et al. (2010) suggested that soluble tau species, rather than fibrillar tau, may be the critical toxic moiety underlying neurodegeneration.

Amino-terminally truncated, pyroglutamylated (pE) forms of amyloid-beta (104760) are strongly associated with Alzheimer disease, are more toxic than amyloid-beta(1-42) and amyloid-beta(1-40), and have been proposed as initiators of Alzheimer disease pathogenesis. Nussbaum et al. (2012) reported a mechanism by which pE-amyloid-beta may trigger Alzheimer disease. Amyloid-beta-3(pE)-42 co-oligomerizes with excess amyloid-beta(1-42) to form metastable low-n oligomers (LNOs) that are structurally distinct and far more cytotoxic to cultured neurons than comparable LNOs made from amyloid-beta(1-42) alone. Tau is required for cytotoxicity, and LNOs comprising 5% amyloid-beta-3(pE)-42 plus 95% amyloid-beta(1-42) (5% pE-amyloid-beta) seed new cytotoxic LNOs through multiple serial dilutions into amyloid-beta(1-42) monomers in the absence of additional amyloid-beta-3(pE)-42. LNOs isolated from human Alzheimer disease brain contained amyloid-beta-3(pE)-42, and enhanced amyloid-beta-3(pE)-42 formation in mice triggered neuron loss and gliosis at 3 months, but not in a tau-null background. Nussbaum et al. (2012) concluded that amyloid-beta-3(pE)-42 confers tau-dependent neuronal death and causes template-induced misfolding of amyloid-beta(1-42) into structurally distinct LNOs that propagate by a prion-like mechanism. Nussbaum et al. (2012) concluded that their results raised the possibility that amyloid-beta-3(pE)-42 acts similarly at a primary step in Alzheimer disease pathogenesis.

Using in situ hybridization and immunohistochemical analyses, Lund et al. (2013) showed that TTBK1 (619415) was a neuron-specific kinase in human brain, with primary expression in the somatodendritic compartment. TTBK1 phosphorylated tau at ser422 and colocalized with phosphorylated ser422 pretangle neurons, but not with late-stage neurofibrillary tangles.

Kondo et al. (2015) investigated how traumatic brain injury (TBI), an environmental risk factor for Alzheimer disease, leads to tauopathy. Kondo et al. (2015) found robust cis phosphorylated tau protein (P-tau) pathology after TBI in humans and mice. After TBI in mice and stress in vitro, neurons acutely produced cis P-tau, which disrupted axonal microtubule networks and mitochondrial transport, spread to other neurons, and led to apoptosis. This process, which they termed 'cistauosis,' appears long before other tauopathy. Treating TBI mice with cis antibody blocked cistauosis, prevented tauopathy development and spread, and restored many TBI-related structural and functional sequelae. Thus, Kondo et al. (2015) concluded that cis P-tau is a major early driver of disease after TBI and leads to tauopathy in traumatic encephalopathy and Alzheimer disease. The authors suggested that the cis antibody may be further developed to detect and treat TBI and prevent progressive neurodegeneration after injury.

Faraco et al. (2019) reported that dietary salt induced hyperphosphorylation of tau followed by cognitive dysfunction in mice, and that these effects were prevented by restoring endothelial nitric oxide production. The nitric oxide deficiency reduced neuronal calpain (see 114220) nitrosylation and resulted in enzyme activation, which, in turn, led to tau phosphorylation by activating cyclin-dependent kinase-5 (CDK5; 123831). Salt-induced cognitive impairment was not observed in tau-null mice or in mice treated with anti-tau antibodies, despite persistent cerebral hypoperfusion and neurovascular dysfunction. Faraco et al. (2019) concluded that these findings identified a causal link between dietary salt, endothelial dysfunction, and tau pathology, independent of hemodynamic insufficiency. They further suggested that avoidance of excessive salt intake and maintenance of vascular health may help to stave off the vascular and neurodegenerative pathologies that underlie dementia in the elderly.

Rauch et al. (2020) showed that the low density lipoprotein receptor-related protein-1 (LRP1; 107770) controls the endocytosis of tau and its subsequent spread. Knockdown of LRP1 significantly reduced tau uptake in H4 neuroglioma cells and in induced pluripotent stem cell-derived neurons. The interaction between tau and LRP1 is mediated by lysine residues in the microtubule-binding repeat region of tau. Furthermore, downregulation of LRP1 in an in vivo mouse model of tau spread was found to effectively reduce the propagation of tau between neurons. Rauch et al. (2020) concluded that their results identified LRP1 as a key regulator of tau spread in the brain.


Biochemical Features

Cryoelectron Microscopy

Using cryoelectron microscopy, Falcon et al. (2018) determined the structures of tau filaments from patients with Pick disease (172700), a neurodegenerative disorder characterized by frontotemporal dementia. The filaments consist of residues lys254-phe378 of 3R tau, which are folded differently from the tau filaments in Alzheimer disease, establishing the existence of conformers of assembled tau. The observed tau fold in the filaments of patients with Pick disease explains the selective incorporation of 3R tau in Pick bodies, and the differences in phosphorylation relative to the tau filaments of Alzheimer disease. Falcon et al. (2018) concluded that their findings showed how tau can adopt distinct folds in the human brain in different diseases, an essential step for understanding the formation and propagation of molecular conformers.

Using cryoelectron microscopy, Falcon et al. (2019) determined the structure of tau filaments from the brains of 3 individuals with chronic traumatic encephalopathy (CTE) at resolutions down to 2.3 angstroms. Falcon et al. (2019) showed that filament structures were identical in the 3 cases but were distinct from those of Alzheimer disease and Pick disease, and from those formed in vitro. Similar to Alzheimer disease, all 6 brain tau isoforms assemble into filaments in CTE, and residues K274-R379 of 3-repeat tau and S305-R379 of 4-repeat tau form the ordered core of 2 identical C-shaped protofilaments. However, a different conformation of the beta-helix region creates a hydrophobic cavity that is absent in tau filaments from the brains of patients with Alzheimer disease. This cavity encloses an additional density that is not connected to tau, which suggests that the incorporation of cofactors may have a role in tau aggregation in CTE. Moreover, filaments in CTE have distinct protofilament interfaces to those of Alzheimer disease. Falcon et al. (2019) concluded that their structures provided a unifying neuropathologic criterion for CTE, and supported the hypothesis that the formation and propagation of distinct conformers of assembled tau underlie different neurodegenerative diseases.

Using cryoelectron microscopy, Zhang et al. (2020) analyzed the structures of tau filaments extracted from the brains of 3 individuals with corticobasal degeneration (CBD). These filaments were identical between cases, but distinct from those seen in Alzheimer disease (104300), Pick disease (172700), and chronic traumatic encephalopathy. The core of a CBD filament comprises residues lysine-274 to glutamate-380 of tau, spanning the last residue of the R1 repeat, the whole of the R2, R3, and R4 repeats, and 12 amino acids after R4. The core adopts a previously unseen 4-layered fold, which encloses a large nonproteinaceous density. This density is surrounded by the side chains of lysine residues 290 and 294 from R2 and lysine-370 from the sequence after R4.


Population Genetics

The MAPT gene has a polymorphic inversion resulting in 2 main haplotypes: the H1 haplotype, which comprises the more common human noninverted sequence, and the H2 haplotype, which comprises an inverted sequence and includes several other genes within its 900-kb extent (review by Donnelly et al., 2010). The promoter in H1 chromosomes is more efficient at driving transcription than the promoter sequence on the H2 haplotype (Kwok et al., 2004).

Studies by Baker et al. (1999), Skipper et al. (2004), and others identified extensive linkage disequilibrium in the 17q21.31 region and considerable divergence between the H1 and H2 lineages of MAPT. Using a refined physical map of chromosome 17q21.31, Stefansson et al. (2005) uncovered a 900-kb inversion polymorphism in this region accounting for these observations. Chromosomes with the inverted segment in different orientations represent the 2 distinct lineages H1 and H2 that have diverged for as much as 3 million years and show no evidence of having recombined. The H2 lineage is rare in Africans, almost absent in East Asians, but is found at a rate of approximately 20% in Europeans, in whom the haplotype structure is indicative of a history of positive selection. The frequency is about 6% in Africans and less than 1% in East Asians. Stefansson et al. (2005) showed that the H2 lineage is undergoing positive selection in the Icelandic population, such that carrier females have more children and have higher recombination rates than noncarriers.

Zody et al. (2008) used comparative sequencing approaches to investigate the evolutionary history of the European-enriched 17q21.31 MAPT inversion polymorphism. The authors presented a detailed BAC-based sequence assembly of the inverted human H2 haplotype and compared it to the sequence structure and genetic variation of the corresponding 1.5-Mb region for the noninverted H1 human haplotype and that of chimpanzee and orangutan. Zody et al. (2008) found that the inversion of the MAPT region is similarly polymorphic in other great ape species, and presented evidence that the inversions occurred independently in chimpanzees and humans. In humans, the inversion breakpoints correspond to core duplications with the LRRC37 gene family (see 616555). The analysis of Zody et al. (2008) favored the H2 configuration and sequence haplotype as the likely great ape and human ancestral state, with inversion recurrences during primate evolution. The authors further showed that the H2 architecture has evolved more extensive sequence homology, perhaps explaining its tendency to undergo microdeletion associated with mental retardation in European populations.

Donnelly et al. (2010) genotyped SNPs in the MAPT gene corresponding to the H1 and H2 haplotypes in 3,135 individuals from 66 populations worldwide. The H2 inversion was found at the highest frequencies in Southwest Asia and Southern Europe (approximately 30%). Elsewhere in Europe, the frequency of H2 varied from less than 5% in Finns to 28% in Orcadians. The H2 inversion haplotype occurs at low frequencies (1 to 10%) in Africa, Central Asia, East Asia, and the Americas, although the East Asian and Amerindian alleles may be due to recent gene flow from Europe. Molecular evolution analyses indicated that the H2 haplotype originally arose in Africa or Southwest Asia. The most recent common human ancestor was estimated to be from 13,600 to 108,400 years ago, which is much more recent than the 3 million year age estimated by Stefansson et al. (2005).


Molecular Genetics

Rademakers et al. (2004) reported that in the previous 5 years, research had identified 34 different pathogenic MAPT mutations in 101 families worldwide. They described the considerable differences in clinical and pathologic presentation of patients with MAPT mutations and summarized the effect of the different mutations on tau functioning. In addition, they discussed the role of tau as a genetic susceptibility factor, together with the genetic evidence for additional causal genes for tau-positive as well tau-negative dementia.

Frontotemporal Dementia with Parkinsonism

In 13 families with autosomal dominant frontotemporal dementia with parkinsonism linked to chromosome 17 (FTDP17; 600274), Hutton et al. (1998) identified mutations in the MAPT gene: 3 were missense mutations (157140.0001-157140.0003) and 3 were mutations in the 5-prime splice site of exon 10 (157140.0004-157140.0006). All of the splice site mutations destabilized a potential stem-loop structure likely involved in regulating the alternative splicing of exon 10 (Goedert et al., 1989), resulting in more frequent usage of the 5-prime splice site and an increased proportion of tau transcripts that included exon 10. The increase in exon 10+ mRNA was expected to increase the proportion of tau containing 4 microtubule-binding repeats, consistent with the neuropathology described in families with FTDP17. Most cases showed neuronal and/or glial inclusions that stained positively with antibodies raised against tau, although the tau pathology varied considerably in both its quantity and characteristics; the disorder is one of several termed 'tauopathies.'

Varani et al. (1999) determined the 3-dimensional structure of the tau exon 10 splicing regulatory element RNA by means of NMR spectroscopy. They showed that it does indeed form a stable, folded stem-loop structure whose thermodynamic stability is reduced by FTDP17 mutations and increased by compensatory mutations. By exon trapping, Varani et al. (1999) showed that the reduction in thermodynamic stability was correlated with increased splicing of exon 10.

Hong et al. (1998) indicated that more than 10 exonic and intronic mutations of the MAPT gene had been identified in about 20 FTDP17 families. Analyses of soluble and insoluble tau proteins from brains of FTDP17 patients indicated that different pathogenic mutations differentially altered distinct biochemical properties and stoichiometry of brain tau isoforms. Functional assays of recombinant tau proteins with different FTDP17 missense mutations implicated all but 1 of these mutations in disease pathogenesis by reducing the ability of tau to bind microtubules and promote microtubule assembly.

In a study of frontotemporal dementia in the Netherlands from January 1994 to June 1998, Rizzu et al. (1999) found 37 patients who had one or more first-degree relatives with dementia. A mutation in the MAPT gene was found in 17.8% of the group of patients with FTDP17 and in 43% of patients with FTDP17 who also had a positive family history of the disorder. Three distinct missense mutations, G272V (157140.0002), P301L (157140.0001), and R406W (157140.0003) accounted for 15.6% of the mutations. The missense mutations and a single amino acid deletion that was detected in 1 patient, strongly reduced the ability of tau to promote microtubule assembly. Rizzu et al. (1999) suggested that the MAPT mutations cause disturbances in the interactions of the tau protein with microtubules, resulting in hyperphosphorylation of tau protein, assembly into filaments, and subsequent cell death.

Verpillat et al. (2002) found that the tau H1/H1 genotype (see below) was significantly overrepresented in 100 patients with frontotemporal dementia compared to controls (odds ratio for H1/H1 = 1.95). In addition, there was a significant negative effect in carriers of both the H1/H1 genotype and the APOE2 allele (107741).

Goedert et al. (1998) reviewed the role of tau mutations in frontotemporal dementias. Heutink (2000) reviewed the role of tau protein in frontotemporal dementia and other neurodegenerative disorders. Hutton (2001) reviewed the known missense and splice site mutations in the tau gene that are associated with disease and described different mechanisms involved in pathogenesis, including disruption of the interaction between tau and tubulin, deposition of abnormal tau filaments, and the generation of abnormal ratios of tau isoforms.

Using purified recombinant proteins, Alonso et al. (2004) showed that several FTDP17-associated tau missense mutations made tau a more favorable substrate for abnormal hyperphosphorylation compared with wildtype tau. Both the phosphorylation kinetics, due to induced conformational changes, and the phosphorylation stoichiometry, due to increased phosphorylation of more than a single site, were more favorable in the mutant proteins. The mutant proteins polymerized into filaments more readily than wildtype tau, leading to decreased ability to bind wildtype tau.

Progressive Supranuclear Palsy

Conrad et al. (1997) demonstrated an association between progressive supranuclear palsy (PSNP; 601104) and a dinucleotide TG repeat polymorphism in intron 9 of the MAPT gene. They demonstrated overrepresentation of the most common allele (a0) and genotype (a0/a0) in PSNP. Baker et al. (1999) identified a series of polymorphisms scattered throughout the MAPT gene and described 2 extended haplotypes, designated H1 and H2, that cover the entire gene. The dinucleotide TG polymorphism alleles a0 (11 repeats), a1 (12 repeats), and a2 (13 repeats) are inherited with the H1 haplotype, whereas the a3 (14 repeats) and a4 (15 repeats) alleles are inherited with the H2 haplotype. In a total of approximately 200 unrelated Caucasian individuals, there was complete disequilibrium between polymorphisms that spanned the gene, suggesting that the establishment of the 2 haplotypes was an ancient event and that either recombination was suppressed in this region, or recombinant genes were selected against. Baker et al. (1999) showed that the more common haplotype, designated H1, was significantly overrepresented in patients with progressive supranuclear palsy, extending earlier reports of the association between the intronic dinucleotide polymorphism allele a0 and the disorder.

Litvan et al. (2001) examined 63 patients with PSNP and found that the presence of the tau H1/H1 genotype was significantly greater in patients compared to controls. There was no difference between PSNP cases with one H1 or two H1 alleles in the age of onset, severity, or survival of patients, thus showing that tau genotyping does not predict the prognosis of PSNP. However, Litvan et al. (2001) noted that most of the PSNP patients carried the H1/H1 genotype (88.9%) and none of the patients carried the H2/H2 genotype, thus limiting the conclusions of the study.

Using single-nucleotide polymorphisms, Pittman et al. (2004) mapped linkage disequilibrium (LD) in the regions flanking MAPT and established the maximum extent of the haplotype block on chromosome 17q21.31 as a region covering approximately 2 Mb. The gene-rich region extended centromerically beyond the corticotropin-releasing hormone receptor-1 gene (CRHR1; 122561) to a region of approximately 400 kb, where there was a complete loss of LD. The telomeric end was defined by an approximately 150-kb region just beyond the WNT3 (165330) gene. The authors showed that the entire, fully extended H1 haplotype was associated with PSNP, which implicates several other genes in addition to MAPT as candidate pathogenic loci.

Rademakers et al. (2005) and Pittman et al. (2005) used a large collection of pathologically confirmed PSNP samples to fine map PSNP risk on H1 chromosomes in PSNP cases and controls. PSNP risk was associated with an extended subhaplotype (H1c), and the risk for PSNP was narrowed to a 22-kb region in intron 0 of MAPT by examining younger patients with, presumably, a larger genetic component to their disease. The most likely explanation of the association of the MAPT H1 haplotype and PSNP is that variants in the H1 (and H2) haplotypes confer risk of (protect against) disease by altering expression at the locus, with the risky H1 haplotype expressing higher levels of MAPT.

Kwok et al. (2008) showed that the MAPT G-to-A allele of rs242557, which partially defines the H1c subhaplotype, results in increased MAPT gene expression.

Pick Disease

Zhukareva et al. (2002) used biochemical, immunohistochemical, and ultrastructural methods to characterize pathologic tau isoform composition in 14 sporadic Pick disease (172700) brains. They found that both 3R and 4R microtubule-binding isoforms were present in gray and white matter of various brain regions, particularly the cortex and hippocampus. Specifically, 7 cases had predominantly pathologic 3R isoforms, 4 cases had a mixture of 3R and 4R isoforms, and 3 cases had primarily 4R isoforms. Isolated tau filaments were primarily straight, but twisted forms were also present. Although the cases shared similar clinical and neuropathologic features, the biochemical profiles of abnormal tau were diverse.

Parkinson Disease

In a large study of 1,056 individuals from 235 families selected from 13 clinical centers in the United States and Australia and from a family ascertainment core center, Martin et al. (2001) found that haplotypes of single-nucleotide polymorphisms (SNPs) yielded strong evidence of association with late-onset Parkinson disease (PD; 168600). Positive association was found with 1 haplotype (P = 0.009) and a negative association with another haplotype (P = 0.007). The results were thought to implicate MAPT as a susceptibility gene for idiopathic Parkinson disease.

Kwok et al. (2004) identified several SNPs in the MAPT promoter region corresponding to the H1 and H2 haplotypes, as well as a novel variant of the H1 haplotype, termed H1-prime. In a cohort of 206 patients with idiopathic late-onset Parkinson disease, the authors found a significant association with the H1/H1 promoter genotype. In vitro analysis showed that the H1 haplotype was more efficient at driving gene expression than the H2 haplotype, suggesting that increased MAPT expression is a susceptibility factor in idiopathic PD.

Alternate mRNA splicing of exons 2, 3, and 10 of the MAPT gene results in the expression of 6 polypeptides in the human CNS (Higuchi et al., 2002). The predominant isoforms differ by the presence of either 3 or 4 microtubule-binding domains, the 3-repeat (3R) and 4-repeat (4R) isoforms, which result from the exclusion or inclusion of exon 10 (Panda et al., 2003). Frontotemporal dementia with parkinsonism linked to chromosome 17 (FTDP17; 600274) is called, in pathology terms, a '4R tauopathy' because of the presence of fibrillar aggregates comprising the 4R tau isoform. There are 2 predominant MAPT haplotypes--termed H1 and H2 and extending more than 500 kb--in which variants appear to be in complete linkage disequilibrium; H1 and H2 haplotypes do not recombine (Pastor et al., 2002). H1/H1 homozygous genotypes are overrepresented in 4R tauopathies. Using H1-specific SNPs, Skipper et al. (2004) demonstrated that MAPT H1 is a misnomer and consists of a family of recombining H1 alleles. Population genetics, linkage disequilibrium, and association analyses showed that specific MAPT H1 subhaplotypes are preferentially associated with Parkinson disease. Using a sliding scale of MAPT H1-specific haplotypes in age- and sex-matched PD cases and controls from central Norway, Skipper et al. (2004) refined the disease association to within an interval of approximately 90 kb of the 5-prime end of the MAPT locus.

In a study of 557 PD patient-control pairs, Mamah et al. (2005) found that individuals with the SNCA Rep1 261/261 or MAPT H1/H1 genotypes had an increased risk of PD compared to those with neither genotype (odds ratio of 1.96); however, the combined effect of the 2 genotypes was the same as for either genotype alone. Mamah et al. (2005) suggested that the MAPT H1/H1 genotype may cause increased SNCA fibrillization in persons with lower SNCA protein concentrations due to genotypes other than Rep1 261/261. In persons with the Rep1 261/261 genotype, the MAPT H1/H1 genotype confers no additional risk because the SNCA protein is already at threshold concentration for self-fibrillization.

Kwok et al. (2005) identified 2 functional SNPs in the GSK3B (605004) gene that influenced its transcriptional activity and correlated with enhanced phosphorylation of MAPT in vitro, respectively. Conditional logistic regression analysis of the genotypes of 302 Caucasian PD patients and 184 Chinese PD patients found an association between the GSK3B polymorphisms, MAPT haplotype, and risk of PD. Kwok et al. (2005) concluded that GSK3B polymorphisms interact with MAPT haplotypes to modify disease risk in PD. Garcia-Gorostiaga et al. (2009) confirmed the findings of Kwok et al. (2005) in a cohort of 314 Spanish patients with PD.

Among 1,762 PD patients, Zabetian et al. (2007) found a significant association between the H1/H1 genotype and risk of disease (odds ratio of 1.46; p = 8 x 10(-7)). The effect was evident in both familial and sporadic subgroups, men and women, and early- and late-onset disease. Within H1/H1 individuals, there was no association with H1 subhaplotypes.

Among 659 PD patients, Goris et al. (2007) found a synergistic interaction between the MAPT H1 haplotype and an A-to-G SNP (rs356219) in the 3-prime region of the SNCA gene. Carrying the combination of risk genotypes at both loci approximately doubled the risk of disease (p = 3 x 10(-6)). The findings suggested that MAPT and SNCA are involved in shared or converging pathogenic pathways and may have a synergistic effect. Cognitive decline and the development of dementia was associated with the H1/H1 genotype (p = 10(-4)). In a final analysis that combined data from other studies, Goris et al. (2007) confirmed the association of the H1/H1 genotype with PD (odds ratio of 1.4; p = 2 x 10(-19)).

In a study of 543 PD patients from 296 families with a proband and at least 1 affected first-degree relative as part of the GenePD Study group, Tobin et al. (2008) found a significant association between the H1 haplotype and PD (odds ratio of 1.72; p = 0.0008). In particular, rs1800547 of the H1 haplotype was significantly associated with PD (p = 0.02 after Bonferroni correction). Tobin et al. (2008) also identified a novel H1 subhaplotype that predicted an even greater increased risk for PD (OR, 4.48; p = 0.003), but some of these SNPs extended beyond the 3-prime region of MAPT. The expression of 4R MAPT, STH (607067), and KIAA1267 was significantly increased in cerebellum samples from PD patients relative to controls. No difference in expression was observed for 3R repeat MAPT. The findings suggested that variations in MAPT gene expression may underlie the association with PD.

Among 202 Spanish patients with PD, Seto-Salvia et al. (2011) found a significant association between the H1 haplotype and the development of dementia (odds ratio of 3.73, p = 0.002). Examination of subhaplotypes showed that a rare version of H1, named H1p, was overrepresented in PD patients with dementia compared to controls (2.3% vs 0.1%, p = 0.003). There was a protective effect for the H2a haplotype on PD with dementia. There was no association between MAPT variants and disease among 164 patients with Alzheimer disease or 41 with Lewy body dementia.

In a statistical analysis of 5,302 PD patients and 4,161 controls from 15 sites, Elbaz et al. (2011) found no evidence for an interactive effect between the H1 haplotype in the MAPT gene and SNPs in the SNCA gene on disease. Variation in each gene was associated with PD risk, indicating independent effects.

Alzheimer Disease

Conrad et al. (2002) identified a single-nucleotide polymorphism that results in a gly7-to-arg (G7R) change in the saitohin gene which appeared to be overrepresented in the homozygous state in late-onset Alzheimer disease subjects.

Because the MAPT R406W mutation (157140.0003) can cause a clinical picture closely resembling Alzheimer disease (see 104300) and quite different from frontotemporal dementia with parkinsonism, Rademakers et al. (2003) stated that MAPT should be considered a candidate gene for clinical Alzheimer disease families in which mutations in known Alzheimer disease genes have been excluded.

Myers et al. (2005) reported that the H1c subhaplotype of MAPT on the background of the well-described H1 clade was associated with risk of Alzheimer disease in 360 autopsy-confirmed cases with ages at death over 65 years of age and 252 controls.

Kwok et al. (2008) showed that the MAPT G-to-A allele of rs242557, which partially defines the H1c subhaplotype, results in increased MAPT gene expression. The authors also provided evidence that the H1/H2 MAPT haplotype interacts with functional SNPs in the GSK3B gene (605004) to affect risk of Alzheimer disease.


Genotype/Phenotype Correlations

Among 22 patients with FTLD (600274) due to a MAPT mutation, Whitwell et al. (2009) found different patterns of gray matter atrophy using MRI voxel-based morphometry. All patients showed gray matter loss in the anterior temporal lobes, with varying degrees of involvement of the frontal and parietal lobes. Within the temporal lobe, individuals with the IVS10+16 (157140.0006), IVS10+3, N279K (157140.0009), or S305N (157140.0010) mutations showed gray matter loss particularly affecting the medial temporal lobes, including the hippocampus and amygdala. These mutations are all predicted to influence the alternative splicing of MAPT pre-mRNA, resulting in increased 4R tau isoforms. In contrast, patients with the P301L (157140.0001) or V337M (157140.0008) mutations showed gray matter loss particularly affecting the inferior and lateral temporal lobes, with a relative sparing of the medial temporal lobe. P301L and V337M mutation carriers also showed gray matter loss in the basal ganglia. These mutations are predicted to affect the structure and functional properties of the tau protein, which are more prone to aggregation. The different patterns suggested a potential difference in mutant protein function resulting from different pathogenic mutations.


Animal Model

Stambolic et al. (1996) observed that lithium treatment of intact eukaryotic cells inhibits GSK3-dependent phosphorylation of the microtubule-associated protein tau, a putative GSK3 substrate.

Using Western blot analysis, Hiesberger et al. (1999) demonstrated that mice lacking either Reelin (RELN; 600514) or Vldlr (192977) and ApoER2 (LRP8; 602600) exhibit a dramatic increase in the phosphorylation level of the microtubule-stabilizing protein tau.

To model tauopathies, Ishihara et al. (1999) overexpressed the smallest human tau isoform in the central nervous system of transgenic mice. These mice acquired age-dependent central nervous system pathology similar to FTDP17, including insoluble, hyperphosphorylated tau and argyrophilic intraneuronal inclusions formed by tau-immunoreactive filaments. Inclusions were present in cortical and brainstem neurons but were most abundant in spinal cord neurons, where they were associated with axon degeneration, diminished microtubules, and reduced axonal transport in ventral roots, as well as spinal cord gliosis and motor weakness. These transgenic mice recapitulated key features of tauopathies and provided models for elucidating mechanisms underlying diverse tauopathies, including Alzheimer disease.

To model aspects of tau-related physiology and pathology, Spittaels et al. (1999) generated transgenic mice that overexpressed the 4R human tau isoform specifically in neurons. The mice developed axonal degeneration in brain and spinal cord, characterized by reduced or blocked axonal transport and axonal dilations as well as sensorimotor deficits, in the absence of intraneuronal neurofibrillary tangles. Spittaels et al. (1999) noted that excess normal tau might be sufficient to cause neuronal injury and suggested that excess of the 4R tau protein interferes with kinesin-dependent transport by saturating binding sites on microtubules. Spittaels et al. (2000) showed that when GSK3-beta (GSK3B; 605004) is expressed in these transgenic mice there is a strong reduction in the number of axonal dilations and a nearly complete alleviation of motor impairments, presumably by phosphorylation of tau, which then reduces the binding of tau to microtubules. Although more phosphorylated tau was available, neither PHFs nor tangles were formed.

Tesseur et al. (2000) observed that overexpression of human APOE4 in neurons of transgenic mice resulted in hyperphosphorylation of tau, and Tesseur et al. (2000) demonstrated that these mice exhibited widespread astrogliosis in the brain, impaired axonal transport, and axonal degeneration.

Lewis et al. (2000) demonstrated that expression of human tau containing the most common mutation, P301L (157140.0001), results in motor and behavioral deficits in transgenic mice, with age- and gene-dose-dependent development of neurofibrillary tangles (NFT). This phenotype occurred as early as 6.5 months in hemizygous and 4.5 months in homozygous animals. Abnormalities were seen not only in the central nervous system; a peripheral neuropathy and skeletal muscle involvement with neurogenic atrophy were also found.

Gotz et al. (2001) demonstrated that injection of beta-amyloid (104760) A-beta-42 fibrils into the brains of P301L mutant tau transgenic mice causes 5-fold increases in the numbers of neurofibrillary tangles in cell bodies within the amygdala from where neurons project to the injection sites. Gallyas silver impregnation identified neurofibrillary tangles that contained tau phosphorylated at serine-212/threonine-214 and serine-422. Neurofibrillary tangles were composed of twisted filaments and occurred in 6-month-old mice as early as 18 days after A-beta-42 injections. Gotz et al. (2001) concluded that their data support the hypothesis that A-beta-42 fibrils can accelerate neurofibrillary tangle formation in vivo.

Lewis et al. (2001) crossed JNPL3 transgenic mice expressing a mutant tau protein, which developed neurofibrillary tangles and progressive motor disturbance, with Tg2576 transgenic mice expressing mutant beta-amyloid precursor protein (APP), thus modulating the APP-A-beta environment. The resulting double-mutant (tau/APP) progeny and the Tg2576 parental strain developed amyloid-beta deposits at the same age; however, relative to JNPL3 mice, the double mutants exhibited neurofibrillary tangle pathology that was substantially enhanced in the limbic system and olfactory cortex. Lewis et al. (2001) concluded that either APP or amyloid-beta influences the formation of neurofibrillary tangles. The interaction between amyloid-beta and tau pathologies in these mice supports the hypothesis that a similar interaction occurs in Alzheimer disease.

Nguyen et al. (2001) found that hyperphosphorylation of neurofilament and tau proteins was associated with abnormal elevation of the p25/p35 (see 603460) ratio and Cdk5 (123831) activity in the spinal cord of a transgenic mouse model of amyotrophic lateral sclerosis (ALS; 105400).

Using proteomic analysis, David et al. (2005) showed that expression of human P301L mutant tau in transgenic mice resulted in distinct modifications of the brain proteome, suggesting alterations in the mitochondrial electron transport chain, cellular antioxidant capacities, and synaptic properties. Subsequent examination of complex V levels in brains of FTDP17 patients carrying the P301L tau mutation confirmed the observations made in P301L tau transgenic mice and suggested that P301L mutant tau pathology caused a specific mitochondrial dysfunction in humans and mice. In agreement, transgenic P301L tau mice exhibited an initial defect in mitochondrial function with reduced complex I activity, which, with age, translated into a mitochondrial respiration deficiency with diminished ATP synthesis corresponding to reduced complex V activity. P301L mutant tau also caused higher oxidative stress, modified lipid peroxidation levels, and upregulated antioxidant enzyme activities, without reducing mitochondrial numbers or significantly changing transport of mitochondria along neurites. In addition, P301L mutant tau decreased the membrane potential of cortical brain cells in transgenic mice, as these cells became more susceptible to A-beta treatment.

Taniguchi et al. (2005) found that transgenic mice expressing human N279K (157140.0009) mutant tau were viable, with normal feeding and body weight. However, transgenic mice displayed cognitive/sensorimotor deficits when evaluated by a comprehensive battery of tests.

SantaCruz et al. (2005) found that mice expressing a repressible human tau variant developed progressive age-related neurofibrillary tangles, neuronal loss, and behavioral impairments. After the suppression of transgenic tau, memory function recovered and neuron numbers stabilized, but neurofibrillary tangles continued to accumulate. SantaCruz et al. (2005) concluded that neurofibrillary tangles are not sufficient to cause cognitive decline or neuronal death in this model of tauopathy.

Karsten et al. (2006) identified the Npepps gene (606793) as a tau modifier by using a cross-species functional genomic approach to analyze gene expression in mice. Npepps expression was increased in multiple brain regions in a mouse model of frontotemporal dementia (FTD) compared to control mice. In Drosophila, Npepps protected against tau-induced neurodegeneration, whereas loss of Npepps exacerbated neurodegeneration. Immunoblot, SDS-PAGE, and Western blot analyses showed that human NPEPPS directly proteolyzed and significantly diminished human tau. Western blot analysis of 6 brains derived from human FTD patients showed increased NPEPPS expression, particularly in the cerebellum.

Yoshiyama et al. (2007) developed transgenic mice expressing wildtype human MAPT or MAPT with the pro301-to-ser (P301S; 157140.0012) mutation. P301S mice developed synaptic pathology and microgliosis in hippocampus at 3 months of age, followed by synaptic dysfunction at 6 months of age, prior to neuron loss and formation of neurofibrillary tangles. Yoshiyama et al. (2007) concluded that synaptic pathology and microgliosis may be the earliest manifestation of tauopathies.

Chatterjee et al. (2009) employed a Drosophila model of tauopathy to investigate the interdependence of tau kinases in regulating the phosphorylation and toxicity of tau in vivo. Tau mutants resistant to phosphorylation by Par1, the fly homolog of MARK1 (606511), were less toxic than wildtype tau; however, this was not due to their resistance to phosphorylation by Shaggy (GSK3B; 605004). On the contrary, a tau mutant resistant to phosphorylation by Shaggy retained substantial toxicity and had increased affinity for microtubules compared with wildtype tau. Chatterjee et al. (2009) suggested that, in addition to tau phosphorylation, microtubule binding may play a crucial role in the regulation of tau toxicity when misexpressed.

Expression of human tau in C. elegans neurons causes accumulation of aggregated tau leading to neurodegeneration and uncoordinated movement. Guthrie et al. (2009) used this model of human tauopathy disorders to screen for genes required for tau neurotoxicity. Recessive loss-of-function mutations in the Sut2 locus (ZC3H14; 613279) suppress the worm Unc phenotype, tau aggregation, and neurodegenerative changes caused by human tau. Guthrie et al. (2009) cloned the Sut2 gene and found that it encodes a novel subtype of CCCH zinc finger protein conserved across animal phyla. Sut2 shares significant identity with the mammalian SUT2. A yeast 2-hybrid screen revealed that Sut2 bound to Zyg12, the sole C. elegans HOOK protein family member (see 607820). Likewise, Sut2 bound Zyg12 in in vitro protein binding assays. Loss of Zyg12 led to a marked upregulation of Sut2 protein supporting the connection between Sut2 and Zyg12. The human ortholog of Sut2 bound only to HOOK2 (607824), suggesting that the interaction between Sut2 and HOOK family proteins may be conserved across animal phyla.

Iijima-Ando et al. (2010) showed that the DNA damage-activated checkpoint kinase-2 (CHK2; 604373) is a novel tau kinase. Overexpression of Drosophila Chk2 increased tau phosphorylation at ser262 and enhanced tau-induced neurodegeneration in transgenic flies expressing human tau. The nonphosphorylatable ser262-to-ala mutation abolished Chk2-induced enhancement of tau toxicity, suggesting that the ser262 phosphorylation site may be involved in the enhancement of tau toxicity by Chk2. In vitro kinase assays revealed that human CHK2 and a closely related checkpoint kinase, CHK1 (603078), directly phosphorylated human tau at ser262. Drosophila Chk2 did not modulate the activity of the fly homolog of microtubule affinity regulating kinase (see MARK3, 602678), which has been shown to be a physiologic tau ser262 kinase. Iijima-Ando et al. (2010) suggested that CHK1 and CHK2 may be involved in tau phosphorylation and toxicity in the pathogenesis of Alzheimer disease.

Using transgenic Drosophila expressing human A-beta-42 (APP; 104760) and tau, Iijima et al. (2010) showed that tau phosphorylation at ser262 plays a critical role in A-beta-42-induced tau toxicity. Coexpression of A-beta-42 increased tau phosphorylation at AD-related sites including ser262, and enhanced tau-induced neurodegeneration. In contrast, formation of either sarkosyl-insoluble tau or paired helical filaments was not induced by A-beta-42. Coexpression of A-beta-42 and tau carrying the nonphosphorylatable ser262ala mutation did not cause neurodegeneration, suggesting that the ser262 phosphorylation site is required for the pathogenic interaction between A-beta-42 and tau. CHK2 phosphorylates tau at ser262 and enhances tau toxicity in a transgenic Drosophila model (Iijima-Ando et al., 2010). Exacerbation of A-beta-42-induced neuronal dysfunction by blocking tumor suppressor p53 (191170), a key transcription factor for the induction of DNA repair genes, in neurons suggested that induction of a DNA repair response is protective against A-beta-42 toxicity. The authors concluded that tau phosphorylation at ser262 is crucial for A-beta-42-induced tau toxicity in vivo, and they suggested a model of AD progression in which activation of DNA repair pathways is protective against A-beta-42 toxicity but may trigger tau phosphorylation and toxicity in AD pathogenesis.

Using Drosophila and mouse models of tauopathies, Falzone et al. (2010) showed that reductions in axonal transport by reduction in kinesin-1 (KLC1; 600025) expression can exacerbate human tau protein hyperphosphorylation, formation of insoluble aggregates, and tau-dependent neurodegeneration. The authors hypothesized that nonlethal reductions in axonal transport, and perhaps other types of minor axonal stress, are sufficient to induce and/or accelerate abnormal tau behavior characteristic of Alzheimer disease and other neurodegenerative tauopathies.

Xu et al. (2010) generated transgenic mice overexpressing full-length human TTBK1 and tau with the P301L mutation. Transgenic mice developed age-dependent pathology in central nervous system, including intraneuronal accumulation of phosphorylated tau, accumulation of sarkosyl-soluble phospho-tau multimers, Cdk5 and Gsk3-beta activation, locomotor dysfunction, and motor neuron degeneration. The results suggested that TTBK1 is involved in phosphorylation-dependent generation of pathogenic tau aggregation.

Lei et al. (2012) found decreased levels of soluble tau and increased iron levels in the substantia nigra of postmortem samples from patients with Parkinson disease compared to controls. The tau loss was independent of neuronal loss. Similar changes were observed in the MPTP mouse model of Parkinson disease. Mapt-knockout mice developed age-dependent brain atrophy, iron accumulation in various brain regions, and neuronal loss in the substantia nigra, as well as cognitive deficits and parkinsonism. These changes could be prevented by oral treatment with a moderate iron chelator, clioquinol. In primary neuronal cell cultures, Lei et al. (2012) found that loss of tau caused iron retention by decreasing the surface expression of APP, which interacts with ferroportin (SLC40A1; 604653) to accomplish neuronal iron efflux. The findings suggested that tau is needed to prevent age-related damage, that loss of soluble tau contributes to toxic neuronal iron accumulation in Alzheimer disease, Parkinson disease, and other tauopathies, and that pharmacologic intervention in this process may ameliorate the disease.

Taylor et al. (2018) showed that transgenic C. elegans expressing the kinase catalytic domain of human TTBK1 or TTBK2 (611695) were behaviorally normal. However, C. elegans coexpressing TTBK1 or TTBK2 with tau caused neurodegeneration, behavioral abnormalities, aberrant phosphorylation, and shortened lifespan. Coexpression of TTBK2 with high levels of tau resulted in lethality.

Przybyla et al. (2020) found that transgenic mice expressing human tau with a pro301-to-ser (P301S) mutation displayed progressive spatial learning deficits accompanied by reduced synaptic plasticity and aberrant neuronal network activity in brain. Mutant mice presented with an immediate early gene response, consistent with increased neuronal activity. Increased immediate early gene activity was confined to neurons harboring tau pathology, providing a cellular link between aberrant tau and network dysfunction.


History

Neve et al. (1986) noted that 1 of their tau clones hybridized to both chromosome 17 and to an additional region of homology on chromosome 6q21 (MAPTL). They suggested that either the 2 clones represented 2 different tau genes on different chromosomes, or that the 2 clones encoded different regions of the same expressed gene on chromosome 17 and only one of them bore homology to a nonexpressed pseudogene on chromosome 6.

Volz et al. (1994) reported that MAPTL clustered with FTHP1 and GSTA1 (138359) in the region 6p21.1-p12, with the map order unknown.


ALLELIC VARIANTS ( 25 Selected Examples):

.0001 DEMENTIA, FRONTOTEMPORAL, WITH PARKINSONISM

SUPRANUCLEAR PALSY, PROGRESSIVE, 1, INCLUDED
MAPT, PRO301LEU
  
RCV000015313...

In a large Dutch kindred with frontotemporal dementia (600274) reported by Heutink et al. (1997), Hutton et al. (1998) found a pro301-to-leu mutation (P301L) in exon 10 of the MAPT gene. The same mutation was found in a small kindred from the United States. This substitution occurred in a highly conserved region of the MAPT sequence, where a proline residue was found in all mammalian species from which tau had been cloned to that time. The P301L mutation would affect only the 4-repeat tau isoforms because exon 10 is spliced out of mRNA that encodes the 3-repeat isoforms. Analysis of tau aggregates in affected brains from the U.S. kindred revealed that these consist mainly of 4-repeat isoforms, consistent with the mutation affecting exon 10.

In a note added in proof, Poorkaj et al. (1998) described finding a C-to-T transition at nucleotide 728 of the MAPT gene in a newly ascertained family with FTDP17. The mutation resulted in a PRO243LEU mutation. This is the same mutation as that reported by Hutton et al. (1998), who used a different numbering system for the nucleotides and codons. At the time the paper of Poorkaj et al. (1998) was submitted, the longest amino acid form of tau in the database did not include all of the alternatively spliced exons (Poorkaj, 1998).

MAPT transcripts that contain this exon 10 mutation encode tau isoforms with 4 microtubule (MT)-binding repeats (4Rtau) as opposed to tau isoforms with 3 MT-binding repeats (3Rtau). Clark et al. (1998) found that brains of patients with the P301L missense mutation contained aggregates of insoluble 4Rtau in filamentous inclusions, which may lead to neurodegeneration.

Using purified recombinant proteins, Alonso et al. (2004) showed that several FTDP17-associated tau mutations, including P301L, made tau a more favorable substrate for abnormal hyperphosphorylation compared with wildtype tau. Both the phosphorylation kinetics, due to induced conformational changes, and the phosphorylation stoichiometry, due to increased phosphorylation of more than a single site, were more favorable in the mutant proteins. The mutant proteins polymerized into filaments more readily than wildtype tau, leading to decreased ability to bind wildtype tau.

Using proteomic analysis, David et al. (2005) showed that expression of human P301L mutant tau in transgenic mice resulted in distinct modifications of the brain proteome, suggesting alterations in the mitochondrial electron transport chain, cellular antioxidant capacities, and synaptic properties. Subsequent examination of complex V levels in brains of FTDP17 patients carrying the P301L tau mutation confirmed the observations made in P301L tau transgenic mice and suggested that P301L mutant tau pathology caused a specific mitochondrial dysfunction in humans and mice. In agreement, transgenic P301L tau mice exhibited an initial defect in mitochondrial function with reduced complex I activity, which, with age, translated into a mitochondrial respiration deficiency with diminished ATP synthesis corresponding to reduced complex V activity. P301L mutant tau also caused higher oxidative stress, modified lipid peroxidation levels, and upregulated antioxidant enzyme activities, without reducing mitochondrial numbers or significantly changing transport of mitochondria along neurites. In addition, P301L mutant tau decreased the membrane potential of cortical brain cells in transgenic mice, as these cells became more susceptible to A-beta treatment.

Using purified recombinant proteins, Aoyagi et al. (2007) showed that P301L mutant tau assembled into nuclei more rapidly than wildtype tau or R406W (157140.0003) mutant tau. However, P301L mutant nuclei could only promote assembly of P301L mutant tau into filaments, whereas wildtype and R406W mutant nuclei had the ability to seed both wildtype and P301L mutant tau. Pronase digestion experiments revealed conformational differences between P301L mutant tau and wildtype or R406W mutant tau. The core structure of P301L mutant tau seeds was distinct from that of wildtype tau seeds, regardless of phosphorylation state, whereas R406W mutant tau seeds had a core structure similar to that of wildtype tau seeds.

Donker Kaat et al. (2009) identified a P301L mutation in 1 of 172 probands with progressive supranuclear palsy (601104).


.0002 DEMENTIA, FRONTOTEMPORAL, WITH PARKINSONISM

MAPT, GLY272VAL
  
RCV000015315...

In a second large Dutch kindred with frontotemporal dementia (600274) reported by Heutink et al. (1997) as an example of hereditary Pick disease, Hutton et al. (1998) found a gly272-to-val mutation (G272V) that affected a highly conserved residue within the microtubule-associated domain, encoded by exon 9 of the MAPT gene.

Using purified recombinant proteins, Alonso et al. (2004) showed that several FTDP17-associated tau mutations, including G272V, made tau a more favorable substrate for abnormal hyperphosphorylation compared with wildtype tau. Both the phosphorylation kinetics, due to induced conformational changes, and the phosphorylation stoichiometry, due to increased phosphorylation of more than a single site, were more favorable in the mutant proteins. The mutant proteins polymerized into filaments more readily than wildtype tau, leading to decreased ability to bind wildtype tau.


.0003 DEMENTIA, FRONTOTEMPORAL

MAPT, ARG406TRP
  
RCV000015316...

In affected members of a family from the United States with autosomal dominant frontotemporal dementia (600274) originally reported by Reed et al. (1997), Hutton et al. (1998) identified a heterozygous arg406-to-trp mutation (R406W) in the MAPT gene. Neuropathologic examination identified tau-positive intraneuronal neurofibrillary tangles similar to those found in Alzheimer disease (see 104300).

Connell et al. (2001) studied the in vitro phosphorylation of several tau mutants by glycogen synthase kinase 3-beta and confirmed previous findings that the R406W mutation had the surprising effect of reducing tau phosphorylation in cells, compared to both the wildtype and other mutant forms. Connell et al. (2001) suggested that reduced tau phosphorylation and a minor loss of function associated with R406W may contribute to the later onset and somewhat milder phenotype found in families with this mutation.

Tatebayashi et al. (2002) showed that the expression of the R406W mutation in transgenic mice resulted in the development of congophilic hyperphosphorylated tau inclusions in forebrain neurons. These inclusions appeared as early as 18 months of age. As with human cases, tau inclusions were composed of both mutant and endogenous wildtype tau, and were associated with microtubule disruption and flame-shaped transformations of the affected neurons. Behaviorally, aged transgenic mice had associative memory impairment without obvious sensorimotor deficits. Therefore, these mice exhibited a phenotype mimicking R406W FTDP17 (frontotemporal dementia and parkinsonism linked to chromosome 17).

Saito et al. (2002) reported a patient who presented at age 47 years with psychiatric disturbances, primarily delusions, who developed overt dementia by age 52, and died at age 53. There was rapid progression in the last year of life. His father had had a similar illness. Postmortem examination revealed neuronal loss associated with neurofibrillary tangles and neuropil threads, accentuated in the medial temporal lobe, that were immunoreactive for the tau protein. Molecular analysis revealed an R406W mutation.

Rademakers et al. (2003) suggested that the R406W mutation can cause a clinical picture closely resembling Alzheimer disease and quite different from frontotemporal dementia with parkinsonism. They described a 6-generation Belgian family that carried the R406W mutation and had such clinical features, and pointed to reports of 2 other families that carried the R406W mutation and had similar clinical characteristics: 1 from the U.S. Midwest with a Danish ancestor (Reed et al., 1997) and 1 from the Netherlands (van Swieten et al., 1999). Haplotype data ruled against a founder effect for the origin of the mutation in western Europe. Rademakers et al. (2003) stated that their report illustrated the phenotypic heterogeneity of MAPT mutations and reemphasized that MAPT should be considered a candidate gene for clinical Alzheimer disease families in which mutations in known Alzheimer disease genes have been excluded.

Using purified recombinant proteins, Alonso et al. (2004) showed that several FTDP17-associated tau mutations, including R406W, made tau a more favorable substrate for abnormal hyperphosphorylation compared with wildtype tau. Both the phosphorylation kinetics, due to induced conformational changes, and the phosphorylation stoichiometry, due to increased phosphorylation of more than a single site, were more favorable in the mutant proteins. The mutant proteins polymerized into filaments more readily than wildtype tau, leading to decreased ability to bind wildtype tau.

Using purified recombinant proteins, Aoyagi et al. (2007) showed that P301L (157140.0001) mutant tau assembled into nuclei more rapidly than wildtype tau or R406W mutant tau. However, P301L mutant nuclei could only promote assembly of P301L mutant tau into filaments, whereas wildtype and R406W mutant nuclei had the ability to seed both wildtype and P301L mutant tau. Pronase digestion experiments revealed conformational differences between P301L mutant tau and wildtype or R406W mutant tau. The core structure of P301L mutant tau seeds was distinct from that of wildtype tau seeds, regardless of phosphorylation state, whereas R406W mutant tau seeds had a core structure similar to that of wildtype tau seeds.


.0004 DEMENTIA, FRONTOTEMPORAL, WITH PARKINSONISM

MAPT, IVS10, C-U, +14
  
RCV000015317...

In addition to 3 missense mutations, Hutton et al. (1998) found 3 heterozygous mutations in a cluster of 4 nucleotides 13- to 16-bp 3-prime of the exon 10 5-prime splice site of the MAPT gene in families with frontotemporal dementia with parkinsonism (600274). Six families had mutations at 1 of these 3 sites, including 4 families in which the disorder had previously been linked to chromosome 17. One of these linked families was reported by Wilhelmsen et al. (1994) as disinhibition-dementia-parkinsonism-amyotrophy complex (DDPAC; 600274). In this family, the mRNA showed a C-to-U transition in the stem-loop structure at position 14 in the splice donor site of intron 10.


.0005 DEMENTIA, FRONTOTEMPORAL, WITH PARKINSONISM

MAPT, IVS10, A-G, +13
  
RCV000015318...

One of the splice site mutations identified by Hutton et al. (1998) in families with frontotemporal dementia with parkinsonism (600274) was an A-to-G transition at position 13 in intron 10.

In a British patient with frontotemporal dementia, Pickering-Brown et al. (2002) identified the tau IVS10 +13 mutation. He presented with apathy and inertia, and later developed semantic loss.


.0006 DEMENTIA, FRONTOTEMPORAL, WITH PARKINSONISM

MAPT, IVS10, C-T, +16
  
RCV000084537...

In an Australian family with frontotemporal dementia with parkinsonism (600274), Hutton et al. (1998) found a C-to-T change in the MAPT mRNA at position 16 of the splice donor site of intron 10. The mutation resulted in an increased incorporation of exon 10 in MAPT mRNA levels, which increased the proportion of tau isoforms containing 4 microtubule-binding domains.

Goedert et al. (1999) identified a heterozygous IVS10+16C-T mutation in the MAPT gene in affected members of a family with frontotemporal dementia and extrapyramidal signs. Initial neuropathologic examination (Lanska et al., 1994) showed prominent subcortical gliosis (221820), but later studies by Goedert et al. (1999) showed hyperphosphorylated tau in both neurons and glial cells with wide twisted ribbons made of 4-repeat tau isoforms. The molecular findings confirmed the diagnosis. Goedert et al. (1999) noted that phenotypic heterogeneity associated with MAPT mutations has led to classification of related diseases into distinct entities.

Janssen et al. (2002) described the clinical characteristics of 9 families with frontotemporal dementia and the tau exon 10 +16 mutation and found considerable variation in age at onset and duration of disease both between and within families, suggesting the influence of other genetic or environmental factors.

Pickering-Brown et al. (2002) reported 8 British families with frontotemporal dementia from North Wales in which affected members had the tau exon 10 +16 mutation. Clinical features included disinhibition, restless overactivity, fatuous affect, puerile behavior, verbal and motor stereotypes, and semantic loss. In 5 of these 8 families, Pickering-Brown et al. (2004) found a common 3-cM haplotype flanked by markers D17S1860 and D17S806. An affected Australian pedigree (Dark, 1997), 7 patients from London (Janssen et al., 2002; Lantos et al., 2002), and 2 patients from Philadelphia (Poorkaj et al., 2001) had the same haplotype. The disease haplotype was not observed in 100 geographically matched control chromosomes. Pickering-Brown et al. (2004) concluded that the exon 10 +16 mutation represents a founder effect that originated in North Wales. They demonstrated that the mutation is on the H1 tau haplotype.

Doran et al. (2007) reported a large family from Liverpool, England, in which 8 individuals had frontotemporal dementia associated with the IVS10+16C-T mutation. All patients were initially diagnosed with Alzheimer disease (104300) because of presentation of memory deficits and word-finding difficulties. Prototypic features of frontotemporal dementia, such as disinhibition and personality changes, were not noted initially. Doran et al. (2007) noted the phenotypic variability of this mutation.

Colombo et al. (2009) noted that the identification of the IVS10+16C-T mutation in patients from North America and the U.K. support the hypothesis of a founder effect of British origin. Using several different methods of haplotype analysis, the authors estimated that the mutation occurred about 23 generations ago, around 1300 A.D., before Welsh people started emigrating to the U.S. and Australia, where they introduced the mutation.


.0007 DEMENTIA, FRONTOTEMPORAL, WITH PARKINSONISM

MAPT, IVS10, G-A, +1
  
RCV000015320

In a kindred known to have presenile dementia identified as multiple system tauopathy with presenile dementia (MSTD; 600274) in individuals in 5 successive generations, Spillantini et al. (1998) found a G-to-A transition in the nucleotide 3-prime of the exon 10 splice donor site. It was identified in 11 affected family members and segregated with the disease haplotype in other family members. Examination of the nucleotide sequence of exon 10 and the 5-prime intron junction identified a predicted stem-loop structure that encompassed the last 6 nucleotides at the 3-prime end of exon 10 and 19 nucleotides of the intron, including the GT splice donor site. The G-to-A transition destabilized this stem-loop structure. The disorder was associated with an abnormal preponderance of soluble tau protein isoforms with 4 microtubule-binding repeats over isoforms with 3 repeats. This was thought most likely to account for the previous finding that sarkosyl-insoluble tau protein extracted from the filamentous deposits in familial MSTD consists only of tau isoforms with 4 repeats. The departure from the normal ratio of 4-repeat to 3-repeat tau isoforms leads to the formation of abnormal tau filaments. The dysregulation of tau protein production in turn leads to neurodegeneration.


.0008 DEMENTIA, FRONTOTEMPORAL, WITH PARKINSONISM

MAPT, VAL337MET
  
RCV000015321...

Poorkaj et al. (1998) identified 9 DNA sequence variants in 2 families with what they referred to as FTDP17 (600274); 8 of these were also identified in controls and were thus considered polymorphisms. The ninth variant, VAL279MET, was found in 1 FTDP17 family, but not in the other. (This mutation is designated val337 to met in the numbering used by Hutton et al. (1998); see 157140.0001.) The change was considered positive since it occurred in a highly conserved residue and a normal valine is found at this position in all 3 tau interrepeat sequences and in other MAPT homologs. Furthermore, the mutation cosegregated with the disease in the family and was not found in normal controls. At the time that the paper of Poorkaj et al. (1998) was submitted, the longest amino acid form of tau in the database did not include all of the alternatively spliced exons (Poorkaj, 1998).

Using purified recombinant proteins, Alonso et al. (2004) showed that several FTDP17-associated tau mutations, including V337M, made tau a more favorable substrate for abnormal hyperphosphorylation compared with wildtype tau. Both the phosphorylation kinetics, due to induced conformational changes, and the phosphorylation stoichiometry, due to increased phosphorylation of more than a single site, were more favorable in the mutant proteins. The mutant proteins polymerized into filaments more readily than wildtype tau, leading to decreased ability to bind wildtype tau.

Sohn et al. (2019) found that the FTD-associated tau V337M mutant shortened the axon initial segment (AIS) and impaired AIS plasticity in human induced pluripotent stem cell (iPSC)-derived neurons. Electrophysiologic properties of tau V337M neurons revealed that the mutation also impaired homeostatic control of spontaneous neuronal activity in response to depolarization. End-binding protein-3 (EB3, or MAPRE3; 605788), a component of the AIS cytoskeleton, was associated with ANKG (106280) and tau in the AIS submembrane region of human neurons. However, the V337M mutation increased the binding affinity of tau with EB3, leading to increased EB3 accumulation in the AIS in tau V337M mutant neurons. EB3 accumulation increased its interaction with ANKG and promoted its structural stability via physical interaction with microtubules, thereby impairing AIS plasticity in tau V337M mutant neurons.


.0009 DEMENTIA, FRONTOTEMPORAL, WITH PARKINSONISM

MAPT, ASN279LYS
  
RCV000015322...

In affected members of a kindred with pallidopontonigral degeneration (PPND; 600274) originally reported by Wszolek et al. (1992), Clark et al. (1998) identified an asn279-to-lys (N279K) mutation in exon 10 of the MAPT gene. Although the clinical features and associated regional variations in the neuronal loss observed in different kindreds with frontotemporal dementia and parkinsonism linked to chromosome 17 are diverse, the diagnostic lesions in the brain are tau-rich filaments in the cytoplasm of specific subpopulations of neurons and glial cells. The insoluble tau aggregates isolated from brains of patients with the N279K mutation were analyzed by immunoblotting using tau-specific antibodies. For each of 3 mutations, abnormal tau with an apparent relative mass of 64 and 69 kD was observed. The dephosphorylated material comigrated with tau isoforms containing exon 10 having 4 microtubule-binding repeats but not with 3-repeat tau. Thus, the brains contained aggregates of insoluble 4Rtau in filamentous inclusions, which may lead to neurodegeneration.

Delisle et al. (1999) reported 2 French brothers with the N279K mutation who presented early in the fourth decade with a neurodegenerative disorder characterized by an akinetic rigid syndrome and dementia. There was widespread neuronal and glial tau accumulation in the cortex, basal ganglia, brainstem nuclei, and white matter.

Arima et al. (2000) reported 2 Japanese brothers with the N279K mutation who presented with frontotemporal dementia characterized by personality changes, behavioral disinhibition, asocial misconduct, recent memory deficits, and parkinsonism. The disease progressed in both patients, rendering them mute and bedridden with death at ages 57 and 50. Pathologic examination showed severe temporal lobe atrophy, neuronal loss, and tau-immunoreactive neurofibrillary tangles and cytoplasmic inclusions. There was also phosphorylated tau deposition in neurons of the spinal cord and degeneration of the lateral corticospinal tracts. Insoluble tau, mainly 4-repeat isoforms, with molecular masses of 64 and 69 kD were observed. The tau aggregates were found to be composed of paired hollow tubules, 11 to 12 nm in diameter. One of the patients had responded temporarily to L-DOPA therapy.

Yasuda et al. (1999) reported a Japanese patient with the N279K mutation who presented at 41 years of age with rapidly progressive parkinsonism which later included dementia, hallucinations, vertical gaze palsy, extensor plantar responses, frontal release signs, and incontinence. His sister had parkinsonism and dementia and died in her forties. Neuropathologic examination of the proband showed severe cortical atrophy, discoloration of the striatum, globus pallidus, and luysian body, and prominent depigmentation of the substantia nigra and locus ceruleus. In both cortical and subcortical regions, there was moderate to severe neuronal loss, astrocytic gliosis, tau-positive neuronal inclusion bodies, and loss of motor neurons. The authors termed this disease pallido-nigro-luysian degeneration. Wszolek et al. (2000) suggested that the disorder in the patient reported by Yasuda et al. (1999), in which parkinsonism was the initial prominent feature, is a subtype of FTDP17.

Tsuboi et al. (2002) analyzed clinical and genealogic records of 4 previously reported families with the N279K mutation: an American family (Clark et al., 1998), a French family (Delisle et al., 1999), and 2 Japanese families (Yasuda et al., 1999; Arima et al., 2000). The clinical phenotype was similar in all families: onset in the forties, survival time of 6 to 8 years, parkinsonism as the presenting sign, short or no response to L-DOPA treatment, progressive dementia, pyramidal dysfunction, and gaze palsy. Molecular genetic studies showed a shared disease haplotype between the 2 Japanese families, suggesting a founder effect. Tsuboi et al. (2002) suggested that FTDP17 could be divided into 2 major clinical groups, parkinsonism-predominant and dementia-predominant, and that patients with the N279K mutation fall into the former category. Tsuboi et al. (2002) also reported a previously undescribed Japanese patient with the N279K mutation and a similar phenotype.

Rademakers et al. (2004) likewise commented on the highly similar clinical phenotypes associated with the N279K mutation despite the occurrence on different genetic backgrounds. In the various families, the mean age at onset ranged from 41 to 47 years. In patients, typical parkinsonian features such as bradykinesia, rigidity, and postural instability were uniformly seen. Personality and behavioral changes and dementia also occurred during the course of the illness, but were less prominent.

Taniguchi et al. (2005) found that transgenic mice expressing human N279K mutant tau were viable, with normal feeding and body weight. However, transgenic mice displayed cognitive/sensorimotor deficits when evaluated by a comprehensive battery of tests.


.0010 DEMENTIA, FRONTOTEMPORAL, WITH PARKINSONISM

MAPT, SER305ASN
  
RCV000015323...

Iijima et al. (1999) reported a Japanese family with early-onset hereditary frontotemporal dementia (600274) and a novel ser305-to-asn mutation in the tau gene. The patients presented with personality changes followed by impaired cognition and memory as well as disorientation, but minimal parkinsonism.


.0011 PICK DISEASE

MAPT, GLY389ARG
  
RCV000015324...

Murrell et al. (1999) described a gly389-to-arg (G389R) missense mutation in exon 13 of the MAPT gene in a patient with a condition closely resembling Pick disease (172700). When 38 years old, the proband presented with progressive aphasia and memory disturbance, followed by apathy, indifference, and hyperphagia. MRI showed the dramatic progression of cerebral atrophy. PET revealed marked glucose hypometabolism that was most severe in the left frontal, temporal, and parietal cortical regions. Rigidity, pyramidal signs, and profound dementia progressed until death at 43 years of age. A paternal uncle, who had died at 43 years of age, had presented with similar symptoms. The proband's brain showed numerous tau-immunoreactive Pick body-like inclusions.

Of 30 cases of pathologically confirmed Pick disease, Pickering-Brown et al. (2000) identified 2 mutations in the tau gene in 2 unrelated patients: a G-to-A change, resulting in a G389R substitution, and an A-to-C change, resulting in a lys257-to-thr substitution (157140.0014). The patient with the G389R mutation showed a decline in intellectual ability with forgetfulness, aggression, and a decline in personal hygiene at age 32, which progressed to death by age 37. Pathologic examination showed severe atrophy and neuronal loss in the frontal cortex with many tau-immunoreactive neuronal inclusions. In vitro, the G389R mutation reduced the ability of tau to promote microtubule assembly by 25 to 30%.


.0012 DEMENTIA, FRONTOTEMPORAL, WITH PARKINSONISM

MAPT, PRO301SER
  
RCV000015325...

In affected members of a family with early-onset frontotemporal dementia with parkinsonism (600274) and seizures, Sperfeld et al. (1999) identified a 1137C-T transition in exon 10 of the MAPT gene, resulting in a pro301-to-ser (P301S) substitution.

In a father and son with frontotemporal dementia and corticobasal degeneration, respectively, Bugiani et al. (1999) identified a P301S substitution in the MAPT gene. Both individuals developed rapidly progressive disease in the third decade. Neuropathic examination of the father showed extensive filamentous pathology made of hyperphosphorylated tau protein. Recombinant P301S tau protein showed a greatly reduced ability to promote microtubule assembly.

Yasuda et al. (2000) identified a P301S mutation in the MAPT gene in a Japanese man with onset of parkinsonism at age 37 and rapidly progressive frontotemporal dementia beginning at age 39. Citing earlier reports of mutations at the same position, the authors noted that codon 301 of the tau gene is a hotspot of pathogenic mutations and that the mutations exhibit variable phenotypes (see P301L; 157140.0001).

Lossos et al. (2003) identified the P301S mutation in 3 affected members of a Jewish family of Algerian origin with rapidly progressive frontotemporal dementia with parkinsonism. Disease onset was characterized by personality changes in the late thirties, followed by progressive cognitive and motor deterioration leading to akinetic mutism or death within 3 to 5 years. Werber et al. (2003) identified the P301S mutation in a 39-year-old Jewish woman of Algerian origin with frontotemporal dementia with parkinsonism. Family history revealed 4 affected first-degree relatives, 1 of whom was still alive and also carried the mutation. The families reported by Lossos et al. (2003) and Werber et al. (2003) were members of an extended Israeli-French kindred (Yasuda et al., 2005).

Yasuda et al. (2005) reported another Japanese family in which 6 members had frontotemporal dementia and parkinsonism caused by a heterozygous P301S mutation. Age at disease onset ranged from 28 to 40 years, and clinical findings were characterized by parkinsonism in all patients with dementia reported only in the 3 younger patients who had detailed medical histories. Neuropsychiatric features of these 3 patients included emotional incontinence, euphoria, apathy, and perseveration. One patient had psychotic features. Neuropathologic examination of these 3 patients showed neuronal loss and gliosis most prominent in the substantia nigra, globus pallidus, and subthalamic nucleus associated with neuropil thread-rich, tau-containing lesions. Yasuda et al. (2005) noted the phenotypic variability associated with the P301S mutation.


.0013 DEMENTIA, FRONTOTEMPORAL

MAPT, ASN296ASN
  
RCV000015326...

In a family with frontotemporal dementia (600274), previously reported by Brown et al. (1996) as having corticobasal degeneration, Spillantini et al. (2000) identified a silent mutation in exon 10 of the tau gene, resulting in exon trapping, or the inclusion of multiple copies of exon 10 in the mRNA transcript. This mutation was thought to lead to an abnormal preponderance of soluble tau protein isoforms with 4 microtubule-binding repeats over isoforms with 3 repeats, perhaps by disrupting a cis-acting element in the region of exon 10. Affected members demonstrated progressive dementia beginning in the fifth or sixth decade with extensive tau pathology and frontotemporal atrophy.


.0014 DEMENTIA, FRONTOTEMPORAL, WITH PARKINSONISM

MAPT, GLU342VAL
  
RCV000015327...

In a patient with familial frontotemporal dementia (600274), Lippa et al. (2000) identified a 1025A-T transversion in exon 12 of the tau gene, resulting in a glu342-to-val (E342V) substitution. The authors found that the levels of the 4R0N isoform were increased in all brain regions, whereas the 4RN1 and 4RN2 isoforms were decreased, a previously undescribed pathologic tau profile. Lippa et al. (2000) suggested that exon 12 may influence alternative splicing of other exons within the tau gene. Pathologic examination revealed frontotemporal neuron loss, intracytoplasmic tau aggregates, and paired helical tau filaments.


.0015 PICK DISEASE

MAPT, LYS257THR
  
RCV000015328...

Of 30 cases of pathologically confirmed Pick disease (172700), Pickering-Brown et al. (2000) identified 2 mutations in the tau gene in 2 unrelated patients: an A-to-C change, resulting in a lys257-to-thr substitution, and the previously identified G389R substitution(157140.0011). Tau-immunoreactive Pick bodies and Pick cells were present. In vitro, the K257T mutation reduced the ability of tau to promote microtubule assembly by 70%.


.0016 PICK DISEASE

MAPT, LYS369ILE
  
RCV000015329...

In a patient with Pick disease (172700) characterized clinically by onset at age 52 of rapidly progressing decline of cognitive and behavioral abilities, and pathologically by severe temporal atrophy and the presence of Pick inclusion bodies and Pick cells, Neumann et al. (2001) identified a mutation in the tau gene, resulting in a lys369-to-ile (K339I) substitution in exon 12. The K369I mutation led to a 90% reduction in the rate of microtubule assembly, and the authors suggested that free mutant tau may assemble abnormally, leading to pathologic changes.


.0017 DEMENTIA, FRONTOTEMPORAL, WITH PARKINSONISM

MAPT, ARG5HIS
  
RCV000015330...

In a patient with late-onset (age 75 years) frontotemporal dementia with parkinsonism (600274), Hayashi et al. (2002) identified a G-to-A transition in exon 1 of the MAPT gene, resulting in an arg5-to-his (R5H) substitution. Pathologic examination revealed frontotemporal atrophy, neuronal loss, widespread tau-immunoreactive glial cytoplasmic inclusions, and insoluble tau filaments composed of 4-repeat tau. The mutation reduced the ability of tau to promote microtubule assembly and promoted fibril formation in vitro. The patient had an elderly brother with dementia who died at age 86 years.


.0018 PICK DISEASE

MAPT, SER320PHE
  
RCV000015331...

In a patient who presented with presenile dementia characterized by mild memory problems at age 38 years, which progressed to major memory problems, personality changes, and aphasia at age 47, followed by death at age 53, Rosso et al. (2002) identified a C-to-T transition in exon 11 of the MAPT gene, resulting in a ser320-to-phe (S320F) substitution. Postmortem examination revealed findings consistent with Pick disease (172700), including focal bilateral atrophy of the anterior temporal lobes, extensive tau pathology in the form of Pick-like bodies, and insoluble tau-immunoreactive filaments. In vitro studies showed that the mutation resulted in a markedly reduced ability of tau to promote microtubule assembly.


.0019 SUPRANUCLEAR PALSY, PROGRESSIVE, 1

MAPT, ARG5LEU
  
RCV000084498...

Among 96 patients with progressive supranuclear palsy (601104), Poorkaj et al. (2002) identified 1 patient with a G-to-T mutation in a highly conserved position in exon 1 of the tau gene, resulting in an arg5-to-leu (R5L) substitution. Functional studies showed that the mutation delayed assembly initiation and lowered the mass of microtubules formed, but the assembly rate was increased compared to normal tau. The authors hypothesized a gain-of-function mutation. (Replacement of arginine by histidine at the same position (157170.0017) causes frontotemporal dementia with parkinsonism.)


.0020 DEMENTIA, FRONTOTEMPORAL, WITH PARKINSONISM

MAPT, IVS10, T-C, +11
  
RCV000015333...

In a Japanese patient with onset of FTDP17 (600274) at age 48 years and mental retardation since his first decade, Miyamoto et al. (2001) identified a heterozygous IVS10+11T-C change in the MAPT gene. An older brother, 1 sister, his mother, and grandfather had similar features, including mental retardation, but all had died and thus could not be tested. It was not clear if the mental retardation was related.


.0021 SUPRANUCLEAR PALSY, PROGRESSIVE, 1, ATYPICAL

PARKINSON DISEASE, LATE-ONSET, SUSCEPTIBILITY TO, INCLUDED
MAPT, ASN296DEL
  
RCV000015334...

In a Spanish patient with atypical progressive supranuclear palsy (260540), born from a third-degree consanguineous marriage, Pastor et al. (2001) identified a homozygous 3-bp deletion (AAT) of asparagine at codon 296 in the MAPT gene. The proband and his brother demonstrated a remarkably similar phenotype characterized by onset in the late thirties of mild cognitive decline, inappropriate behavior, ocular movement abnormalities, and asymmetric parkinsonism. The parents were both heterozygous for the mutation, consistent with autosomal recessive inheritance. Two uncles, who were heterozygous for the mutation, developed late-onset typical Parkinson disease (168600). However, there were several asymptomatic heterozygous individuals over the age of 60 in the family, which the authors attributed to reduced penetrance. Pastor et al. (2001) noted that codon 296 lies in the region of the gene which alters splicing of exon 10 (see N296N; 157140.0013).

In a family with a variable neurodegenerative phenotype, including a progressive supranuclear palsy-like syndrome and parkinsonism, Rossi et al. (2004) identified a heterozygous deletion of the last 2 bases of codon 296 and the first base of codon 297, resulting in the deletion of asn296. The proband developed antecollis, dysarthria, postural instability with falls, slowing of ocular movements, and increased deep tendon reflexes at age 36 years, consistent with atypical PSNP. The mutation was also identified in a paternal aunt with typical dopa-responsive Parkinson disease, in 2 asymptomatic sisters of the proband, and in 3 asymptomatic daughters of a deceased paternal uncle who had atypical dopa-unresponsive Parkinson disease with pyramidal signs and cognitive impairment. The authors suggested incomplete penetrance of the disorder.

Oliva and Pastor (2004) noted that although the nucleotide changes reported by Rossi et al. (2004) were unique, the resultant amino acid change, deletion of asn296, is the same as that reported by Pastor et al. (2001); therefore the mutation should be considered the same. Following international recommendations (den Dunnen and Antonarakis, 2000), Oliva and Pastor (2004) proposed that the complete nomenclature for this mutation be designated 'c713-715delATA' at the cDNA level and 'delN296' at the protein level. Oliva and Pastor (2004) noted that functional studies had shown that the delN296 change can decrease the ability of 4-repeat tau to promote microtubule assembly (Yoshida et al., 2002). The authors suggested that heterozygosity for the delN296 mutation may be a risk factor for both a PSNP-like syndrome and Parkinson disease.


.0022 DEMENTIA, FRONTOTEMPORAL

MAPT, LEU266VAL
  
RCV000015336...

In 2 brothers with frontotemporal dementia (600274) characterized by onset of personality changes in the thirties with progression to dementia, Kobayashi et al. (2003) identified a heterozygous C-to-G transversion in exon 9 of the MAPT gene, resulting in a leu266-to-val (L266V) substitution. Their mother reportedly had a similar disease, but was not examined. The mutation was not present in 200 control subjects. In addition, both brothers were homozygotes for the H1 tau haplotype, which has been shown to be overrepresented in patients with FTD compared to controls. Brain neuropathologic examination in one brother showed frontotemporal atrophy with severe neuronal loss and gliosis in both the cortices and the substantia nigra. There were also tau-positive neuronal inclusions and tau-positive astrocytes with stout filaments. Recombinant tau proteins with the L266V mutation were less able to promote microtubule assembly than wildtype tau.


.0023 SUPRANUCLEAR PALSY, PROGRESSIVE, 1

MAPT, SER352LEU
  
RCV000084550...

In 2 sibs from a consanguineous marriage who presented with a form of progressive supranuclear palsy (601104) characterized by fatal respiratory hypoventilation, Nicholl et al. (2003) identified a homozygous 1291C-T transition in exon 12 of the MAPT gene, resulting in a nonconserved ser352-to-leu (S352L) substitution in the N-terminal repeat of the tau protein. The authors called the disorder 'tauopathy and respiratory failure.' The 29-year-old pregnant sister developed dyspnea with stridor, and later had a generalized seizure that left her unconscious. Despite therapy, she died after 9 days. Her 30-year-old brother developed cough syncope, dyspnea, and central apnea, which progressed over 40 months, leading to death at age 34 years. During the illness, he showed slow smooth pursuit and impaired saccades, mild rigidity and bradykinesia, and myoclonic jerks. Neuropathologic examination showed widespread neuronal eosinophilia and pyknosis with gliosis in multiple brain regions, consistent with hypoxic brain damage. There was pervasive tau pathology in neuronal perikarya, neurites, and threads in the gray matter of the hippocampus, thalamus, and pons, but not in the cerebral cortex. Functional studies indicated that the mutated protein showed reduced binding to microtubules as well as increased fibrillization and aggregation. Both sibs carried the H1/H1 haplotype associated with PSP, Nicholl et al. (2003) commented on the unusual apparent autosomal recessive inheritance of this tauopathy. In a review of the role of tau in neurodegenerative diseases, Quadros et al. (2007) classified the disorder in the sibs reported by Nicholl et al. (2003) as PSP.


.0024 DEMENTIA, FRONTOTEMPORAL, WITH PARKINSONISM

MAPT, LYS317MET
  
RCV000015338...

In affected members of 2 families from the Basque country in Spain with a tauopathy best described as frontotemporal dementia with parkinsonism (600274), Zarranz et al. (2005) identified a heterozygous A-to-T transversion in exon 11 of the MAPT gene, resulting in a lys317-to-met (K317M) substitution. Mean age at disease onset was 48 years, characterized by dysarthria and features of parkinsonism. All patients developed parkinsonism and a pyramidal syndrome, and half had amyotrophy. Other variable features included supranuclear gaze palsy, bulbar palsy, and dystonia. Although behavioral changes were not a prominent feature, most patients had frontal signs, and cognitive decline appeared late in the disease. All patients eventually became fully dependent, wheelchair- or bed-bound, tetraplegic, mute, and unable to feed orally. Neuropathologic examination showed frontal and temporal lobe atrophy, severe neuronal loss in the substantia nigra, and diffuse patchy neuronal loss and gliosis with numerous tau- and ubiquitin-positive neurons. Six of 7 spinal cords examined showed severe neuronal loss in the anterior horn and degeneration of the corticospinal tracts, similar to that seen in amyotrophic lateral sclerosis. Two bands of phospho-tau of 68 and 64 kD were observed in brain tissue from 1 patient. Zarranz et al. (2005) emphasized the motor component of the syndrome in these patients. Haplotype analysis suggested a common origin in the 2 families, and genealogic analysis identified a probable common ancestor born in 1782.


.0025 SUPRANUCLEAR PALSY, PROGRESSIVE, 1

MAPT, GLY303VAL
  
RCV000084529...

In a patient with progressive supranuclear palsy (601104), Ros et al. (2005) identified a heterozygous 2095G-T transversion in exon 10 of the MAPT gene, resulting in a gly303-to-val (G303V) substitution. Three asymptomatic family members, who were younger than the average age at disease onset, also carried the mutation. Two deceased affected family members were obligate carriers of the mutation. Protein analysis of tissue from the proband's brain detected hyperphosphorylated tau protein and overexpression of tau isoforms with 4 microtubule-binding repeats. The mutation occurred in a highly conserved residue of the protein and was not identified in 194 control chromosomes.


REFERENCES

  1. Abel, K. J., Boehnke, M., Prahalad, M., Ho, P., Flejter, W. L., Watkins, M., VanderStoep, J., Chandrasekharappa, S. C., Collins, F. S., Glover, T. W., Weber, B. L. A radiation hybrid map of the BRCA1 region of chromosome 17q12-q21. Genomics 17: 632-641, 1993. [PubMed: 8244380, related citations] [Full Text]

  2. Alonso, A. D. C., Grundke-Iqbal, I., Iqbal, K. Alzheimer's disease hyperphosphorylated tau sequesters normal tau into tangles of filaments and disassembles microtubules. Nature Med. 2: 783-787, 1996. [PubMed: 8673924, related citations] [Full Text]

  3. Alonso, A. D., Mederlyova, A., Novak, M., Grundke-Iqbal, I., Iqbal, K. Promotion of hyperphosphorylation by frontotemporal dementia tau mutations. J. Biol. Chem. 279: 34873-34881, 2004. [PubMed: 15190058, related citations] [Full Text]

  4. Andreadis, A., Brown, W. M., Kosik, K. S. Structure and novel exons of the human tau gene. Biochemistry 31: 10626-10633, 1992. [PubMed: 1420178, related citations] [Full Text]

  5. Aoyagi, H., Hasegawa, M., Tamaoka, A. Fibrillogenic nuclei composed of P301L mutant tau induce elongation of P301L tau but not wild-type tau. J. Biol. Chem. 282: 20309-20318, 2007. [PubMed: 17526496, related citations] [Full Text]

  6. Arima, K., Kowalska, A., Hasegawa, M., Mukoyama, M., Watanabe, R., Kawai, M., Takahashi, K., Iwatsubo, T., Tabira, T., Sunohara, N. Two brothers with frontotemporal dementia and parkinsonism with an N279K mutation of the tau gene. Neurology 54: 1787-1795, 2000. [PubMed: 10802785, related citations] [Full Text]

  7. Baker, M., Litvan, I., Houlden, H., Adamson, J., Dickson, D., Perez-Tur, J., Hardy, J., Lynch, T., Bigio, E., Hutton, M. Association of an extended haplotype in the tau gene with progressive supranuclear palsy. Hum. Molec. Genet. 8: 711-715, 1999. [PubMed: 10072441, related citations] [Full Text]

  8. Brown, J., Lantos, P. L., Roques, P., Fidani, L., Rossor, M. N. Familial dementia with swollen achromatic neurons and corticobasal inclusion bodies: a clinical and pathological study. J. Neurol. Sci. 135: 21-30, 1996. [PubMed: 8926492, related citations] [Full Text]

  9. Bugiani, O., Murrell, J. R., Giaccone, G., Hasegawa, M., Ghigo, G., Tabaton, M., Morbin, M., Primavera, A., Carella, F., Solaro, C., Grisoli, M., Savoiardo, M., Spillantini, M. G., Tagliavini, F., Goedert, M., Ghetti, B. Frontotemporal dementia and corticobasal degeneration in a family with a P301S mutation in tau. J. Neuropath. Exp. Neurol. 58: 667-677, 1999. [PubMed: 10374757, related citations] [Full Text]

  10. Chatterjee, S., Sang, T.-K., Lawless, G. M., Jackson, G. R. Dissociation of tau toxicity and phosphorylation: role of GSK-3-beta, MARK and Cdk5 in a Drosophila model. Hum. Molec. Genet. 18: 164-177, 2009. [PubMed: 18930955, images, related citations] [Full Text]

  11. Clark, L. N., Poorkaj, P., Wszolek, Z., Geschwind, D. H., Nasreddine, Z. S., Miller, B., Li, D., Payami, H., Awert, F., Markopoulou, K., Andreadis, A., D'Souza, I., Lee, V. M.-Y., Reed, L., Trojanowski, J. Q., Zhukareva, V., Bird, T., Schellenberg, G., Wilhelmsen, K. C. Pathogenic implications of mutations in the tau gene in pallido-ponto-nigral degeneration and related neurodegenerative disorders linked to chromosome 17. Proc. Nat. Acad. Sci. 95: 13103-13107, 1998. [PubMed: 9789048, images, related citations] [Full Text]

  12. Colombo, R., Tavian, D., Baker, M. C., Richardson, A. M. T., Snowden, J. S., Neary, D., Mann, D. M. A., Pickering-Brown, S. M. Recent origin and spread of a common Welsh MAPT splice mutation causing frontotemporal lobar degeneration. Neurogenetics 10: 313-318, 2009. [PubMed: 19365643, related citations] [Full Text]

  13. Connell, J. W., Gibb, G. M., Betts, J. C., Blackstock, W. P., Gallo, J.-M., Lovestone, S., Hutton, M., Anderton, B. H. Effects of FTDP-17 mutations on the in vitro phosphorylation of tau by glycogen synthase kinase 3-beta identified by mass spectrometry demonstrate certain mutations exert long-range conformational changes. FEBS Lett. 493: 40-44, 2001. [PubMed: 11278002, related citations] [Full Text]

  14. Conrad, C., Andreadis, A., Trojanowski, J. Q., Dickson,D. W., Kang, D., Chen, X., Weiderholt, W., Hansen, L., Masliah, E., Thal, L. J., Katzman, R., Xia, Y., Saitoh, T. Genetic evidence for the involvement of tau in progressive supranuclear palsy. Ann. Neurol. 41: 277-281, 1997. [PubMed: 9029080, related citations] [Full Text]

  15. Conrad, C., Vianna, C., Freeman, M., Davies, P. A polymorphic gene nested within an intron of the tau gene: implications for Alzheimer's disease. Proc. Nat. Acad. Sci. 99: 7751-7756, 2002. [PubMed: 12032355, images, related citations] [Full Text]

  16. Dark, F. A family with autosomal dominant non-Alzheimer's presenile dementia. Aust. New Zeal. J. Psychiat. 31: 139-144, 1997. [PubMed: 9088499, related citations] [Full Text]

  17. David, D. C., Hauptmann, S., Scherping, I., Schuessel, K., Keil, U., Rizzu, P., Ravid, R., Drose, S., Brandt, U., Muller, W. E., Eckert, A., Gotz, J. Proteomic and functional analyses reveal a mitochondrial dysfunction in P301L tau transgenic mice. J. Biol. Chem. 280: 23802-23814, 2005. [PubMed: 15831501, related citations] [Full Text]

  18. de Calignon, A., Fox, L. M., Pitstick, R., Carlson, G. A., Bacskai, B. J., Spires-Jones, T. L., Hyman, B. T. Caspase activation precedes and leads to tangles. Nature 464: 1201-1204, 2010. [PubMed: 20357768, images, related citations] [Full Text]

  19. Delacourte, A., Sergeant, N., Champain, D., Wattez, A., Maurage, C.-A., Lebert, F., Pasquier, F., David, J.-P. Nonoverlapping but synergetic tau and APP pathologies in sporadic Alzheimer's disease. Neurology 59: 398-407, 2002. [PubMed: 12177374, related citations] [Full Text]

  20. Delisle, M.-B., Murrell, J. R., Richardson, R., Trofatter, J. A., Rascol, O., Soulages, X., Mohr, M., Calvas, P., Ghetti, B. A mutation at codon 279 (N279K) in exon 10 of the tau gene causes a tauopathy with dementia and supranuclear palsy. Acta Neuropath. 98: 62-77, 1999. [PubMed: 10412802, related citations] [Full Text]

  21. den Dunnen, J. T., Antonarakis, S. E. Mutation nomenclature extensions and suggestions to describe complex mutations: a discussion. Hum. Mutat. 15: 7-12, 2000. Note: Erratum: Hum. Mutat. 20: 403 only, 2002. [PubMed: 10612815, related citations] [Full Text]

  22. Dixit, R., Ross, J. L., Goldman, Y. E., Holzbaur, E. L. F. Differential regulation of dynein and kinesin motor proteins by tau. Science 319: 1086-1089, 2008. [PubMed: 18202255, images, related citations] [Full Text]

  23. Donker Kaat, L., Boon, A. J. W., Azmani, A., Kamphorst, W., Breteler, M. M. B., Anar, B., Heutink, P., van Swieten, J. C. Familial aggregation of parkinsonism in progressive supranuclear palsy. Neurology 73: 98-105, 2009. [PubMed: 19458322, related citations] [Full Text]

  24. Donlon, T. A., Harris, P., Neve, R. L. Localization of microtubule-associated protein tau (MTBT1) to chromosome 17q21. (Abstract) Cytogenet. Cell Genet. 46: 607, 1987.

  25. Donnelly, M. P., Paschou, P., Grigorenko, E., Gurwitz, D., Mehdi, S. Q., Kajuna, S. L. B., Barta, C., Kungulilo, S., Karoma, N. J., Lu, R.-B., Zhukova, O. V., Kim, J.-J., and 11 others. The distribution and most recent common ancestor of the 17q21 inversion in humans. Am. J. Hum. Genet. 86: 161-171, 2010. [PubMed: 20116045, images, related citations] [Full Text]

  26. Doran, M., du Plessis, D. G., Ghadiali, E. J., Mann, D. M. A., Pickering-Brown, S., Larner, A. J. Familial early-onset dementia with tau intron 10 +16 mutation with clinical features similar to those of Alzheimer disease. Arch. Neurol. 64: 1535-1539, 2007. [PubMed: 17923640, related citations] [Full Text]

  27. Elbaz, A., Ross, O. A., Ioannidis, J. P. A., Soto-Ortolaza, A. I., Moisan, F., Aasly, J., Annesi, G., Bozi, M., Brighina, L., Chartier-Harlin, M.-C., Destee, A., Ferrarese, C., and 29 others. Independent and joint effects of the MAPT and SNCA genes in Parkinson disease. Ann. Neurol. 69: 778-792, 2011. [PubMed: 21391235, images, related citations] [Full Text]

  28. Falcon, B., Zhang, W., Murzin, A. G., Murshudov, G., Garringer, H. J., Vidal, R., Crowther, R. A., Ghetti, B., Scheres, S. H. W., Goedert, M. Structures of filaments from Pick's disease reveal a novel tau protein fold. Nature 561: 137-140, 2018. [PubMed: 30158706, images, related citations] [Full Text]

  29. Falcon, B., Zivanov, J., Zhang, W., Murzin, A. G., Garringer, H. J., Vidal, R., Crowther, R. A., Newell, K. L., Ghetti, B., Goedert, M., Scheres, S. H. W. Novel tau filament fold in chronic traumatic encephalopathy encloses hydrophobic molecules. Nature 568: 420-423, 2019. [PubMed: 30894745, images, related citations] [Full Text]

  30. Falzone, T. L., Gunawardena, S., McCleary, D., Reis, G. F., Goldstein, L. S. B. Kinesin-1 transport reductions enhance human tau hyperphosphorylation, aggregation and neurodegeneration in animal models of tauopathies. Hum. Molec. Genet. 19: 4399-4408, 2010. [PubMed: 20817925, images, related citations] [Full Text]

  31. Faraco, G., Hochrainer, K., Segarra, S. G., Schaeffer, S., Santisteban, M. M., Menon, A., Jiang, H., Holtzman, D. M., Anrather, J., Iadecola, C. Dietary salt promotes cognitive impairment through tau phosphorylation. Nature 574: 686-690, 2019. Note: Erratum: Nature 578: E9, 2019. [PubMed: 31645758, images, related citations] [Full Text]

  32. Garcia-Gorostiaga, I., Sanchez-Juan, P., Mateo, I., Rodriguez-Rodriguez, E., Sanchez-Quintana, C., Curiel del Olmo, S., Vazquez-Higuera, J. L., Berciano, J., Combarros, O., Infante, J. Glycogen synthase kinase-3 and tau genes interact in Parkinson's and Alzheimer's diseases. (Letter) Ann. Neurol. 65: 759-760, 2009. [PubMed: 19557862, related citations] [Full Text]

  33. Giasson, B. I., Forman, M. S., Higuchi, M., Golbe, L. I., Graves, C. L., Kotzbauer, P. T., Trojanowski, J. Q., Lee, V. M.-Y. Initiation and synergistic fibrillization of tau and alpha-synuclein. Science 300: 636-640, 2003. [PubMed: 12714745, related citations] [Full Text]

  34. Goedert, M., Crowther, R. A., Spillantini, M. G. Tau mutations cause frontotemporal dementias. Neuron 21: 955-958, 1998. [PubMed: 9856453, related citations] [Full Text]

  35. Goedert, M., Spillantini, M. G., Crowther, R. A., Chen, S. G., Parchi, P., Tabaton, M., Lanska, D. J., Markesbery, W. R., Wilhelmsen, K. C., Dickson, D. W., Petersen, R. B., Gambetti, P. Tau gene mutation in familial progressive subcortical gliosis. Nature Med. 5: 454-457, 1999. [PubMed: 10202939, related citations] [Full Text]

  36. Goedert, M., Spillantini, M. G., Jakes, R., Rutherford, D., Crowther, R. A. Multiple isoforms of human microtubule-associated protein tau: sequences and localization in neurofibrillary tangles of Alzheimer's disease. Neuron 3: 519-526, 1989. [PubMed: 2484340, related citations] [Full Text]

  37. Goedert, M., Spillantini, M. G., Potier, M. C., Ulrich, J., Crowther, R. A. Cloning and sequencing of the cDNA encoding an isoform of microtubule-associated protein tau containing four tandem repeats: differential expression of tau protein mRNAs in human brain. EMBO J. 8: 393-399, 1989. [PubMed: 2498079, related citations] [Full Text]

  38. Goedert, M., Wischik, C. M., Crowther, R. A., Walker, J. E., Klug, A. Cloning and sequencing of the cDNA encoding a core protein of the paired helical filament of Alzheimer disease: identification as the microtubule-associated protein tau. Proc. Nat. Acad. Sci. 85: 4051-4055, 1988. [PubMed: 3131773, related citations] [Full Text]

  39. Goode, B. L., Chau, M., Denis, P. E., Feinstein, S. C. Structural and functional differences between 3-repeat and 4-repeat tau isoforms: implications for normal tau function and the onset of neurodegenerative disease. J. Biol. Chem. 275: 38182-38189, 2000. [PubMed: 10984497, related citations] [Full Text]

  40. Goris, A., Williams-Gray, C. H., Clark, G. R., Foltynie, T., Lewis, S. J. G., Brown, J., Ban, M., Spillantini, M. G., Compston, A., Burn, D. J., Chinnery, P. F., Barker, R. A., Sawcer, S. J. Tau and alpha-synuclein in susceptibility to, and dementia in, Parkinson's disease. Ann. Neurol. 62: 145-153, 2007. [PubMed: 17683088, related citations] [Full Text]

  41. Gotz, J., Chen, F., van Dorpe, J., Nitsch, R. M. Formation of neurofibrillary tangles in P301L tau transgenic mice induced by A-beta42 fibrils. Science 293: 1491-1495, 2001. [PubMed: 11520988, related citations] [Full Text]

  42. Guo, J.-P., Arai, T., Miklossy, J., McGeer, P. L. Amyloid-beta and tau form soluble complexes that may promote self aggregation of both into the insoluble forms observed in Alzheimer's disease. Proc. Nat. Acad. Sci. 103: 1953-1958, 2006. [PubMed: 16446437, images, related citations] [Full Text]

  43. Guthrie, C. R., Schellenberg, G. D., Kraemer, B. C. SUT-2 potentiates tau-induced neurotoxicity in Caenorhabditis elegans. Hum. Molec. Genet. 18: 1825-1838, 2009. [PubMed: 19273536, images, related citations] [Full Text]

  44. Hayashi, S., Toyoshima, Y., Hasegawa, M., Umeda, Y., Wakabayashi, K., Tokiguchi, S., Iwatsubo, T., Takahashi, H. Late-onset frontotemporal dementia with a novel exon 1 (Arg5His) tau gene mutation. Ann. Neurol. 51: 525-530, 2002. [PubMed: 11921059, related citations] [Full Text]

  45. Heutink, P., Stevens, M., Rizzu, P., Bakker, E., Kros, J. M., Tibben, A., Niermeijer, M. F., van Duijn, C. M., Oostra, B. A., van Swieten, J. C. Hereditary frontotemporal dementia is linked to chromosome 17q21-q22: a genetic and clinicopathological study of three Dutch families. Ann. Neurol. 41: 150-159, 1997. [PubMed: 9029063, related citations] [Full Text]

  46. Heutink, P. Untangling tau-related dementia. Hum. Molec. Genet. 9: 979-986, 2000. [PubMed: 10767321, related citations] [Full Text]

  47. Hiesberger, T., Trommsdorff, M., Howell, B. W., Goffinet, A., Mumby, M. C., Cooper, J. A., Herz, J. Direct binding of reelin to VLDL receptor and apoE receptor 2 induces tyrosine phosphorylation of disabled-1 and modulates tau phosphorylation. Neuron 24: 481-489, 1999. [PubMed: 10571241, related citations] [Full Text]

  48. Higuchi, M., Lee, V. M.-Y., Trojanowski, J. Q. Tau and axonopathy in neurodegenerative disorders. Neuromolec. Med. 2: 131-150, 2002. [PubMed: 12428808, related citations] [Full Text]

  49. Holzer, M., Craxton, M., Jakes, R., Arendt, T., Goedert, M. Tau gene (MAPT) sequence variation among primates. Gene 341: 313-322, 2004. [PubMed: 15474313, related citations] [Full Text]

  50. Hong, M., Zhukareva, V., Vogelsberg-Ragaglia, V., Wszolek, Z., Reed, L., Miller, B. I., Geschwind, D. H., Bird, T. D., McKeel, D., Goate, A., Morris, J. C., Wilhelmsen, K. C., Schellenberg, G. D., Trojanowski, J. Q., Lee, V. M.-Y. Mutation-specific functional impairments in distinct tau isoforms of hereditary FTDP-17. Science 282: 1914-1917, 1998. [PubMed: 9836646, related citations] [Full Text]

  51. Hutton, M., Lendon, C. L., Rizzu, P., Baker, M., Froelich, S., Houlden, H., Pickering-Brown, S., Chakraverty, S., Isaacs, A., Grover, A., Hackett, J., Adamson, J., and 39 others. Association of missense and 5-prime-splice-site mutations in tau with the inherited dementia FTDP-17. Nature 393: 702-705, 1998. [PubMed: 9641683, related citations] [Full Text]

  52. Hutton, M. Missense and splice site mutations in tau associated with FTDP-17: multiple pathogenic mechanisms. Neurology 56 (suppl. 4): S21-S25, 2001. [PubMed: 11402146, related citations] [Full Text]

  53. Iijima, K., Gatt, A., Iijima-Ando, K. Tau ser262 phosphorylation is critical for A-beta-42-induced tau toxicity in a transgenic Drosophila model of Alzheimer's disease. Hum. Molec. Genet. 19: 2947-2957, 2010. [PubMed: 20466736, images, related citations] [Full Text]

  54. Iijima, M., Tabira, T., Poorkaj, P., Schellenberg, G. D., Trojanowski, J. Q., Lee, V. M., Schmidt, M. L., Takahashi, K., Nabika, T., Matsumoto, T., Yamashita, Y., Yoshioka, S., Ishino, H. A distinct familial presenile dementia with a novel missense mutation in the tau gene. Neuroreport 10: 497-501, 1999. [PubMed: 10208578, related citations] [Full Text]

  55. Iijima-Ando, K., Zhao, L., Gatt, A., Shenton, C., Iijima, K. A DNA damage-activated checkpoint kinase phosphorylates tau and enhances tau-induced neurodegeneration. Hum. Molec. Genet. 19: 1930-1938, 2010. [PubMed: 20159774, images, related citations] [Full Text]

  56. Ishihara, T., Hong, M., Zhang, B., Nakagawa, Y., Lee, M. K., Trojanowski, J. Q., Lee, V. M.-Y. Age-dependent emergence and progression of a tauopathy in transgenic mice overexpressing the shortest human tau isoform. Neuron 24: 751-762, 1999. [PubMed: 10595524, related citations] [Full Text]

  57. Janssen, J. C., Warrington, E. K., Morris, H. R., Lantos, P., Brown, J., Revesz, T., Wood, N., Khan, M. N., Cipolotti, L., Fox, N. C., Rossor, M. N. Clinical features of frontotemporal dementia due to the intronic tau 10 +16 mutation. Neurology 58: 1161-1168, 2002. [PubMed: 11971081, related citations] [Full Text]

  58. Jiang, Z., Cote, J., Kwon, J. M., Goate, A. M., Wu, J. Y. Aberrant splicing of tau pre-mRNA caused by intronic mutations associated with the inherited dementia frontotemporal dementia with Parkinsonism linked to chromosome 17. Molec. Cell. Biol. 20: 4036-4048, 2000. Note: Erratum: Molec. Cell. Biol. 20: 5360 only, 2000. [PubMed: 10805746, images, related citations] [Full Text]

  59. Karsten, S. L., Sang, T.-K., Gehman, L. T., Chatterjee, S., Liu, J., Lawless, G. M., Sengupta, S., Berry, R. W., Pomakian, J., Oh, H. S., Schulz, C., Hui, K.-S., Wiedau-Pazos, M., Vinters, H. V., Binder, L. I., Geschwind, D. H., Jackson, G. R. A genomic screen for modifiers of tauopathy identified puromycin-sensitive aminopeptidase as an inhibitor of tau-induced neurodegeneration. Neuron 51: 549-560, 2006. [PubMed: 16950154, related citations] [Full Text]

  60. Kobayashi, T., Ota, S., Tanaka, K., Ito, Y., Hasegawa, M., Umeda, Y., Motoi, Y., Takanashi, M., Yasuhara, M., Anno, M., Mizuno, Y., Mori, H. A novel L266V mutation of the tau gene causes frontotemporal dementia with a unique tau pathology. Ann. Neurol. 53: 133-137, 2003. [PubMed: 12509859, related citations] [Full Text]

  61. Kondo, A., Shahpasand, K., Mannix, R., Qiu, J., Moncaster, J., Chen, C.-H., Yao, Y., Lin, Y.-M., Driver, J. A., Sun, Y., Wei, S., Luo, M.-L., and 10 others. Antibody against early driver of neurodegeneration cis P-tau blocks brain injury and tauopathy. Nature 523: 431-436, 2015. [PubMed: 26176913, images, related citations] [Full Text]

  62. Kwok, J. B. J., Hallupp, M., Loy, C. T., Chan, D. K. Y., Woo, J., Mellick, G. D., Buchanan, D. D., Silburn, P. A., Halliday, G. M., Schofield, P. R. GSK3B polymorphisms alter transcription and splicing in Parkinson's disease. Ann. Neurol. 58: 829-839, 2005. [PubMed: 16315267, related citations] [Full Text]

  63. Kwok, J. B. J., Loy, C. T., Hamilton, G., Lau, E., Hallupp, M., Williams, J., Owen, M. J., Broe, G. A., Tang, N., Lam, L., Powell, J. F., Lovestone, S., Schofield, P. R. Glycogen synthase kinase-3 and tau genes interact in Alzheimer's disease. Ann. Neurol. 64: 446-454, 2008. [PubMed: 18991351, related citations] [Full Text]

  64. Kwok, J. B. J., Teber, E. T., Loy, C., Hallupp, M., Nicholson, G., Mellick, G. D., Buchanan, D. D., Silburn, P. A., Schofield, P. R. Tau haplotypes regulate transcription and are associated with Parkinson's disease. Ann. Neurol. 55: 329-334, 2004. [PubMed: 14991810, related citations] [Full Text]

  65. Lanska, D. J., Currier, R. D., Cohen, M., Gambetti, P., Smith, E. E., Bebin, J., Jackson, J. F., Whitehouse, P. J., Markesbery, W. R. Familial progressive subcortical gliosis. Neurology 44: 1633-1643, 1994. [PubMed: 7936288, related citations] [Full Text]

  66. Lantos, P. L., Cairns, N. J., Khan, M. N., King, A., Revesz, T., Janssen, J. C., Morris, H., Rossor, M. N. Neuropathologic variation in frontotemporal dementia due to the intronic tau 10 +16 mutation. Neurology 58: 1169-1175, 2002. [PubMed: 11971082, related citations] [Full Text]

  67. Lei, P., Ayton, S., Finkelstein, D. I., Spoerri, L., Ciccotosto, G. D., Wright, D. K., Wong, B. X. W., Adlard, P. A., Cherny, R. A., Lam, L. Q., Roberts, B. R., Volitakis, I., Egan, G. F., McLean, C. A., Cappai, R., Duce, J. A., Bush, A. I. Tau deficiency induces parkinsonism with dementia by impairing APP-mediated iron export. Nature Med. 18: 291-295, 2012. [PubMed: 22286308, related citations] [Full Text]

  68. Lewis, J., Dickson, D. W., Lin, W.-L., Chisholm, L., Corral, A., Jones, G., Yen, S.-H., Sahara, N., Skipper, L., Yager, D., Eckman, C., Hardy, J., Hutton, M., McGowan, E. Enhanced neurofibrillary degeneration in transgenic mice expressing mutant tau and APP. Science 293: 1487-1491, 2001. [PubMed: 11520987, related citations] [Full Text]

  69. Lewis, J., McGowan, E., Rockwood, J., Melrose, H., Nacharaju, P., Van Slegtenhorst, M., Gwinn-Hardy, K., Murphy, M. P., Baker, M., Yu, X., Duff, K., Hardy, J., Corral, A., Lin, W.-L., Yen, S.-H., Dickson, D. W., Davies, P., Hutton, M. Neurofibrillary tangles, amyotrophy and progressive motor disturbance in mice expressing mutant (P301L) tau protein. Nature Genet. 25: 402-405, 2000. Note: Erratum: Nature Genet. 26: 127 only, 2000. [PubMed: 10932182, related citations] [Full Text]

  70. Li, H.-L., Wang, H.-H., Liu, S.-J., Deng, Y.-Q., Zhang, Y.-J., Tian, Q., Wang, X.-C., Chen, X.-Q., Yang, Y., Zhang, J.-Y., Wang, Q., Xu, H., Liao, F.-F., Wang, J.-Z. Phosphorylation of tau antagonizes apoptosis by stabilizing beta-catenin, a mechanism involved in Alzheimer's neurodegeneration. Proc. Nat. Acad. Sci. 104: 3591-3596, 2007. [PubMed: 17360687, images, related citations] [Full Text]

  71. Lippa, C. F., Zhukareva, V., Kawarai, T., Uryu, K., Shafiq, M., Nee, L. E., Grafman, J., Liang, Y., St George-Hyslop, P. H., Trojanowski, J. Q., Lee, V. M.-Y. Frontotemporal dementia with novel tau pathology and a glu342val tau mutation. Ann. Neurol. 48: 850-858, 2000. [PubMed: 11117541, related citations]

  72. Litvan, I., Baker, M., Hutton, M. Tau genotype: no effect on onset, symptom severity, or survival in progressive supranuclear palsy. Neurology 57: 138-140, 2001. [PubMed: 11445645, related citations] [Full Text]

  73. Liu, F., Iqbal, K., Grundke-Iqbal, I., Hart, G. W., Gong, C.-X. O-GlcNAcylation regulates phosphorylation of tau: a mechanism involved in Alzheimer's disease. Proc. Nat. Acad. Sci. 101: 10804-10809, 2004. [PubMed: 15249677, images, related citations] [Full Text]

  74. Lossos, A., Reches, A., Gal, A., Newman, J. P., Soffer, D., Gomori, J. M., Boher, M., Ekstein, D., Biran, I., Meiner, Z., Abramsky, O., Rosenmann, H. Frontotemporal dementia and parkinsonism with the P301S tau gene mutation in a Jewish family. J. Neurol. 250: 733-740, 2003. [PubMed: 12796837, related citations] [Full Text]

  75. Lu, P.-J., Wulf, G., Zhou, X. Z., Davies, P., Lu, K. P. The prolyl isomerase Pin1 restores the function of Alzheimer-associated phosphorylated tau protein. Nature 399: 784-788, 1999. [PubMed: 10391244, related citations] [Full Text]

  76. Lund, H., Cowburn, R. F., Gustafsson, E., Stromberg, K., Svensson, A., Dahllund, L., Malinwosky, D., Sunnemark, D. Tau-tubulin kinase 1 expression, phosphorylation and co-localization with phospho-ser422 tau in Alzheimer's disease brain. Brain Path. 23: 378-389, 2013. [PubMed: 23088643, images, related citations] [Full Text]

  77. Mamah, C. E., Lesnick, T. G., Lincoln, S. J., Strain, K. J., de Andrade, M., Bower, J. H., Ahlskog, J. E., Rocca, W. A., Farrer, M. J., Maraganore, D. M. Interaction of alpha-synuclein and tau genotypes in Parkinson's disease. Ann. Neurol. 57: 439-443, 2005. [PubMed: 15732111, related citations] [Full Text]

  78. Martin, E. R., Scott, W. K., Nance, M. A., Watts, R. L., Hubble, J. P., Koller, W. C., Lyons, K., Pahwa, R., Stern, M. B., Colcher, A., Hiner, B. C., Jankovic, J., and 20 others. Association of single-nucleotide polymorphisms of the Tau gene with late-onset Parkinson disease. JAMA 286: 2245-2250, 2001. [PubMed: 11710889, related citations] [Full Text]

  79. Miyamoto, K., Kowalska, A., Hasegawa, M., Tabira, T., Takahashi, K., Araki, W., Akiguchi, I., Ikemoto, A. Familial frontotemporal dementia and parkinsonism with a novel mutation at an intron 10+11-splice site in the tau gene. Ann. Neurol. 50: 117-120, 2001. [PubMed: 11456301, related citations] [Full Text]

  80. Murrell, J. R., Spillantini, M. G., Zolo, P., Guazzelli, M., Smith, M. J., Hasegawa, M., Redi, F., Crowther, R. A., Pietrini, P., Ghetti, B., Goedert, M. Tau gene mutation G389R causes a tauopathy with abundant Pick body-like inclusions and axonal deposits. J. Neuropath. Exp. Neurol. 58: 1207-1226, 1999. [PubMed: 10604746, related citations] [Full Text]

  81. Myers, A. J., Kaleem, M., Marlowe, L., Pittman, A. M., Lees, A. J., Fung, H. C., Duckworth, J., Leung, D., Gibson, A., Morris, C. M., de Silva, R., Hardy, J. The H1c haplotype at the MAPT locus is associated with Alzheimer's disease. Hum. Molec. Genet. 14: 2399-2404, 2005. [PubMed: 16000317, related citations] [Full Text]

  82. Nagase, T., Ishikawa, K., Kikuno, R., Hirosawa, M., Nomura, N., Ohara, O. Prediction of the coding sequences of unidentified human genes. XV. The complete sequences of 100 new cDNA clones from brain which code for large proteins in vitro. DNA Res. 6: 337-345, 1999. [PubMed: 10574462, related citations] [Full Text]

  83. Neumann, M., Schulz-Schaeffer, W., Crowther, R. A., Smith, M. J., Spillantini, M. G., Goedert, M., Kretzschmar, H. A. Pick's disease associated with the novel tau gene mutation K369I. Ann. Neurol. 50: 503-513, 2001. [PubMed: 11601501, related citations] [Full Text]

  84. Neve, R. L., Harris, P., Kosik, K. S., Kurnit, D. M., Donlon, T. A. Identification of cDNA clones for the human microtubule-associated protein tau and chromosomal localization of the genes for tau and microtubule-associated protein 2. Brain Res. 387: 271-280, 1986. [PubMed: 3103857, related citations] [Full Text]

  85. Nguyen, M. D., Lariviere, R. C., Julien, J.-P. Deregulation of Cdk5 in a mouse model of ALS: toxicity alleviated by perikaryal neurofilament inclusions. Neuron 30: 135-147, 2001. [PubMed: 11343650, related citations] [Full Text]

  86. Nicholl, D. J., Greenstone, M. A., Clarke, C. E., Rizzu, P., Crooks, D., Crowe, A., Trojanowski, J. Q., Lee, V. M.-Y., Heutink, P. An English kindred with a novel recessive tauopathy and respiratory failure. Ann. Neurol. 54: 682-686, 2003. [PubMed: 14595660, related citations] [Full Text]

  87. Nussbaum, J. M., Schilling, S., Cynis, H., Silva, A., Swanson, E., Wangsanut, T., Tayler, K., Wiltgen, B., Hatami, A., Ronicke, R., Reymann, K., Hutter-Paier, B., Alexandru, A., Jagla, W., Graubner, S., Glabe, C. G., Demuth, H.-U., Bloom, G. S. Prion-like behaviour and tau-dependent cytotoxicity of pyroglutamylated amyloid-beta. Nature 485: 651-655, 2012. [PubMed: 22660329, images, related citations] [Full Text]

  88. Oliva, R., Pastor, P. Tau gene delN296 mutation, Parkinson's disease, and atypical supranuclear palsy. (Letter) Ann. Neurol. 55: 448-449, 2004. [PubMed: 14991828, related citations] [Full Text]

  89. Panda, D., Samuel, J. C., Massie, M., Feinstein, S. C., Wilson, L. Differential regulation of microtubule dynamics by three- and four-repeat tau: implications for the onset of neurodegenerative disease. Proc. Nat. Acad. Sci. 100: 9548-9553, 2003. [PubMed: 12886013, images, related citations] [Full Text]

  90. Pastor, P., Ezquerra, M., Tolosa, E., Munoz, E., Marti, M. J., Valldeoriola, F., Molinuevo, J. L., Calopa, M., Oliva, R. Further extension of the H1 haplotype associated with progressive supranuclear palsy. Mov. Disord. 17: 550-556, 2002. [PubMed: 12112206, related citations] [Full Text]

  91. Pastor, P., Pastor, E., Carnero, C., Vela, R., Garcia, T., Amer, G., Tolosa, E., Oliva, R. Familial atypical progressive supranuclear palsy associated with homozygosity for the delN296 mutation in the tau gene. Ann. Neurol. 49: 263-267, 2001. [PubMed: 11220749, related citations] [Full Text]

  92. Petrucelli, L., Dickson, D., Kehoe, K., Taylor, J., Snyder, H., Grover, A., De Lucia, M., McGowan, E., Lewis, J., Prihar, G., Kim, J., Dillmann, W. H., Browne, S. E., Hall, A., Voellmy, R., Tsuboi, Y., Dawson, T. M., Wolozin, B., Hardy, J., Hutton, M. CHIP and Hsp70 regulate tau ubiquitination, degradation and aggregation. Hum. Molec. Genet. 13: 703-714, 2004. [PubMed: 14962978, related citations] [Full Text]

  93. Pickering-Brown, S., Baker, M., Bird, T., Trojanowski, J., Lee, V., Morris, H., Rossor, M., Janssen, J. C., Neary, D., Craufurd, D., Richardson, A., Snowden, J., Hardy, J., Mann, D., Hutton, M. Evidence of a founder effect in families with frontotemporal dementia that harbor the tau +16 splice mutation. Am. J. Med. Genet. 125B: 79-82, 2004. [PubMed: 14755449, related citations] [Full Text]

  94. Pickering-Brown, S., Baker, M., Yen, S.-H., Liu, W.-K., Hasegawa, M., Cairns, N., Lantos, P. L., Rossor, M., Iwatsubo, T., Davies, Y., Allsop, D., Furlong, R., Owen, F., Hardy, J., Mann, D., Hutton, M. Pick's disease is associated with mutations in the tau gene. Ann. Neurol. 48: 859-867, 2000. [PubMed: 11117542, related citations]

  95. Pickering-Brown, S. M., Richardson, A. M. T., Snowden, J. S., McDonagh, A. M., Burns, A., Braude, W., Baker, M., Liu, W.-K., Yen, S.-H., Hardy, J., Hutton, M., Davies, Y., Allsop, D., Craufurd, D., Neary, D., Mann, D. M. A. Inherited frontotemporal dementia in nine British families associated with intronic mutations in the tau gene. Brain 125: 732-751, 2002. [PubMed: 11912108, related citations] [Full Text]

  96. Pittman, A. M., Myers, A. J., Abou-Sleiman, P., Fung, H. C., Kaleem, M., Marlowe, L., Duckworth, J., Leung, D., Williams, D., Kilford, L., Thomas, N., Morris, C. M., Dickson, D., Wood, N. W., Hardy, J., Lees, A. J., de Silva, R. Linkage disequilibrium fine mapping and haplotype association of the tau gene in progressive supranuclear palsy and corticobasal degeneration. J. Med. Genet. 42: 837-846, 2005. [PubMed: 15792962, related citations] [Full Text]

  97. Pittman, A. M., Myers, A. J., Duckworth, J., Bryden, L., Hanson, M., Abou-Sleiman, P., Wood, N. W., Hardy, J., Lees, A. J., de Silva, R. The structure of the tau haplotype in controls and in progressive supranuclear palsy. Hum. Molec. Genet. 13: 1267-1274, 2004. [PubMed: 15115761, related citations] [Full Text]

  98. Poorkaj, P., Bird, T. D., Wijsman, E., Nemens, E., Garruto, R. M., Anderson, L., Andreadis, A., Wiederholt, W. C., Raskind, M., Schellenberg, G. D. Tau is a candidate gene for chromosome 17 frontotemporal dementia. Ann. Neurol. 43: 815-825, 1998. Note: Erratum: Ann. Neurol. 44: 428 only, 1998. [PubMed: 9629852, related citations] [Full Text]

  99. Poorkaj, P., Grossman, M., Steinbart, E., Payami, H., Sadovnick, A., Nochlin, D., Tabira, T., Trojanowski, J. Q., Borson, S., Galasko, D., Reich, S., Quinn, B., Schellenberg, G., Bird, T. D. Frequency of tau gene mutations in familial and sporadic cases of non-Alzheimer dementia. Arch. Neurol. 58: 383-387, 2001. [PubMed: 11255441, related citations] [Full Text]

  100. Poorkaj, P., Kas, A., D'Souza, I., Zhou, Y., Pham, Q., Stone, M., Olson, M. V., Schellenberg, G. D. A genomic sequence analysis of the mouse and human microtubule-associated protein tau. Mammalian Genome 12: 700-712, 2001. [PubMed: 11641718, related citations] [Full Text]

  101. Poorkaj, P., Muma, N. A., Zhukareva, V., Cochran, E. J., Shannon, K. M., Hurtig, H., Koller, W. C., Bird, T. D., Trojanowski, J. Q., Lee, V. M.-Y., Schellenberg, G. D. An R5L tau mutation in a subject with a progressive supranuclear palsy phenotype. Ann. Neurol. 52: 511-516, 2002. [PubMed: 12325083, related citations] [Full Text]

  102. Poorkaj, P. Personal Communication. Seattle, Wash. 11/10/1998.

  103. Przybyla, M., van Eersel, J., van Hummel, A., van der Hoven, J., Sabale, M., Harasta, A., Muller, J., Gajwani, M., Prikas, E., Mueller, T., Stevens, C. H., Power, J., Housley, G. D., Karl, T., Kassiou, M., Ke, Y. D., Ittner, A., Ittner, L. M. Onset of hippocampal network aberration and memory deficits in P301S tau mice are associated with an early gene signature. Brain 143: 1889-1904, 2020. [PubMed: 32375177, related citations] [Full Text]

  104. Quadros, A., Weeks, O. I., Ait-Ghezala, G. Role of tau in Alzheimer's dementia and other neurodegenerative diseases. J. Appl. Biomed. 5: 1-12, 2007.

  105. Rademakers, R., Cruts, M., van Broeckhoven, C. The role of tau (MAPT) in frontotemporal dementia and related tauopathies. Hum. Mutat. 24: 277-295, 2004. [PubMed: 15365985, related citations] [Full Text]

  106. Rademakers, R., Dermaut, B., Peeters, K., Cruts, M., Heutink, P., Goate, A., Van Broeckhoven, C. Tau (MAPT) mutation arg406trp presenting clinically with Alzheimer disease does not share a common founder in western Europe. Hum. Mutat. 22: 409-411, 2003. [PubMed: 14517953, related citations] [Full Text]

  107. Rademakers, R., Melquist, S., Cruts, M., Theuns, J., Del-Favero, J., Poorkaj, P., Baker, M., Sleegers, K., Crook, R., De Pooter, T., Bel Kacem, S., Adamson, J., and 15 others. High-density SNP haplotyping suggests altered regulation of tau gene expression in progressive supranuclear palsy. Hum. Molec. Genet. 14: 3281-3292, 2005. [PubMed: 16195395, related citations] [Full Text]

  108. Rapoport, M., Dawson, H. N., Binder, L. I., Vitek, M. P., Ferreira, A. Tau is essential to beta-amyloid-induced neurotoxicity. Proc. Nat. Acad. Sci. 99: 6364-6369, 2002. [PubMed: 11959919, images, related citations] [Full Text]

  109. Rauch, J. N., Luna, G., Guzman, E., Audouard, M., Challis, C., Sibih, YE., Leshuk, C., Hernandez, I., Wegmann, S., Hyman, B. T., Gradinaru, V., Kampmann, M., Kosik, K. S. LRP1 is a master regulator of tau uptake and spread. Nature 580: 381-385, 2020. [PubMed: 32296178, images, related citations] [Full Text]

  110. Reed, L. A., Grabowski, T. J., Schmidt, M. L., Morris, J. C., Goate, A., Solodkin, A., Van Hoesen, G. W., Schelper, R. L., Talbot, C. J., Wragg, M. A., Trojanowski, J. Q. Autosomal dominant dementia with widespread neurofibrillary tangles. Ann. Neurol. 42: 564-572, 1997. [PubMed: 9382467, related citations] [Full Text]

  111. Rizzu, P., Hinkle, D. A., Zhukareva, V., Bonifati, V., Severijnen, L.-A., Martinez, D., Ravid, R., Kamphorst, W., Eberwine, J. H., Lee, V. M.-Y., Trojanowski, J. Q., Heutink, P. DJ-1 colocalizes with tau inclusions: a link between parkinsonism and dementia. Ann. Neurol. 55: 113-118, 2004. [PubMed: 14705119, related citations] [Full Text]

  112. Rizzu, P., Van Swieten, J. C., Joosse, M., Hasegawa, M., Stevens, M., Tibben, A., Niermeijer, M. F., Hillebrand, M., Ravid, R., Oostra, B. A., Goedert, M., van Duijn, C. M., Heutink, P. High prevalence of mutations in the microtubule-associated protein tau in a population study of frontotemporal dementia in the Netherlands. Am. J. Hum. Genet. 64: 414-421, 1999. [PubMed: 9973279, related citations] [Full Text]

  113. Roberson, E. D., Scearce-Levie, K., Palop, J. J., Yan, F., Cheng, I. H., Wu, T., Gerstein, H., Yu, G.-Q., Mucke, L. Reducing endogenous tau ameliorates amyloid beta-induced deficits in an Alzheimer's disease mouse model. Science 316: 750-754, 2007. [PubMed: 17478722, related citations] [Full Text]

  114. Ros, R., Thobois, S., Streichenberger, N., Kopp, N., Sanchez, M. P., Perez, M., Hoenicka, J., Avila, J., Honnorat, J., de Yebenes, J. G. A new mutation of the tau gene, G303V, in early-onset familial progressive supranuclear palsy. Arch. Neurol. 62: 1444-1450, 2005. [PubMed: 16157753, related citations] [Full Text]

  115. Rossi, G., Gasparoli, E., Pasquali, C., Di Fede, G., Testa, D., Albanese, A., Bracco, F., Tagliavini, F. Progressive supranuclear palsy and Parkinson's disease in a family with a new mutation in the tau gene. (Letter) Ann. Neurol. 55: 448 only, 2004. [PubMed: 14991829, related citations] [Full Text]

  116. Rosso, S. M., van Herpen, E., Deelen, W., Kamphorst, W., Severijnen, L.-A., Willemsen, R., Ravid, R., Niermeijer, M. F., Dooijes, D., Smith, M. J., Goedert, M., Heutink, P., van Swieten, J. C. A novel tau mutation, S320F, causes a tauopathy with inclusions similar to those in Pick's disease. Ann. Neurol. 51: 373-376, 2002. [PubMed: 11891833, related citations] [Full Text]

  117. Saito, Y., Geyer, A., Sasaki, R., Kuzuhara, S., Nanba, E., Miyasaka, T., Suzuki, K., Murayama, S. Early-onset, rapidly progressive familial tauopathy with R406W mutation. Neurology 58: 811-813, 2002. [PubMed: 11889249, related citations] [Full Text]

  118. SantaCruz, K., Lewis, J., Spires, T., Paulson, J., Kotilinek, L., Ingelsson, M., Guimaraes, A., DeTure, M., Ramsden, M., McGowan, E., Forster, C., Yue, M., Orne, J., Janus, C., Mariash, A., Kuskowski, M., Hyman, B., Hutton, M., Ashe, K. H. Tau suppression in a neurodegenerative mouse model improves memory function. Science 309: 476-481, 2005. [PubMed: 16020737, images, related citations] [Full Text]

  119. Schneider, A., Biernat, J., von Bergen, M., Mandelkow, E., Mandelkow, E. M. Phosphorylation that detaches tau protein from microtubules (Ser262, Ser214) also protects it against aggregation into Alzheimer paired helical filaments. Biochemistry 38: 3549-3558, 1999. [PubMed: 10090741, related citations] [Full Text]

  120. Seto-Salvia, N., Clarimon, J., Pagonabarraga, J., Pascual-Sedano, B., Campolongo, A., Combarros, O., Mateo, J. I., Regana, D., Martinez-Corral, M., Marquie, M., Alcolea, D., Suarez-Calvet, M., Molina-Porcel, L., Dols, O., Gomez-Isla, T., Blesa, R., Lleo, A., Kulisevsky, J. Dementia risk in Parkinson disease: disentangling the role of MAPT haplotypes. Arch. Neurol. 68: 359-364, 2011. [PubMed: 21403021, related citations] [Full Text]

  121. Skipper, L., Wilkes, K., Toft, M., Baker, M., Lincoln, S., Hulihan, M., Ross, O. A., Hutton, M., Aasly, J., Farrer, M. Linkage disequilibrium and association of MAPT H1 in Parkinson disease. Am. J. Hum. Genet. 75: 669-677, 2004. [PubMed: 15297935, images, related citations] [Full Text]

  122. Sohn, P. D., Huang, C. T.-L., Yan, R., Fan, L., Tracy, T. E., Camargo, C. M., Montgomery, K. M., Arhar, T., Mok, S.-A., Freilich, R., Baik, J., He, M., Gong, S., Roberson, E. D., Karch, C. M., Gestwicki, J. E., Xu, K., Kosik, K. S., Gan, L. Pathogenic tau impairs axon initial segment plasticity and excitability homeostasis. Neuron 104: 458-470, 2019. [PubMed: 31542321, images, related citations] [Full Text]

  123. Sperfeld, A. D., Collatz, M. B., Baier, H., Palmbach, M., Storch, A., Schwarz, J., Tatsch, K., Reske, S., Joosse, M., Heutink, P., Ludolph, A. C. FTDP-17: an early-onset phenotype with parkinsonism and epileptic seizures caused by a novel mutation. Ann. Neurol. 46: 708-715, 1999. [PubMed: 10553987, related citations] [Full Text]

  124. Spillantini, M. G., Murrell, J. R., Goedert, M., Farlow, M. R., Klug, A., Ghetti, B. Mutation in the tau gene in familial multiple system tauopathy with presenile dementia. Proc. Nat. Acad. Sci. 95: 7737-7741, 1998. [PubMed: 9636220, images, related citations] [Full Text]

  125. Spillantini, M. G., Yoshida, H., Rizzini, C., Lantos, P. L., Khan, N., Rossor, M. N., Goedert, M., Brown, J. A novel tau mutation (N296N) in familial dementia with swollen achromatic neurons and corticobasal inclusion bodies. Ann. Neurol. 48: 939-943, 2000. [PubMed: 11117553, related citations] [Full Text]

  126. Spittaels, K., van den Haute, C., van Dorpe, J., Bruynseels, K., Vandezande, K., Laenen, I., Geerts, H., Mercken, M., Sciot, R., van Lommel, A., Loos, R., van Leuven, F. Prominent axonopathy in the brain and spinal cord of transgenic mice overexpressing four-repeat human tau protein. Am. J. Path. 155: 2153-2165, 1999. [PubMed: 10595944, images, related citations] [Full Text]

  127. Spittaels, K., van den Haute, C., van Dorpe, J., Geerts, H., Mercken, M., Bruynseels, K., Lasrado, R., Vandezande, K., Laenen, I., Boon, T., van Lint, J., Vandenheede, J, Moechars, D., Loos, R., van Leuven, F. Glycogen synthase kinase-3-beta phosphorylates protein tau and rescues the axonopathy in the central nervous system of human four-repeat tau transgenic mice. J. Biol. Chem. 275: 41340-41349, 2000. [PubMed: 11007782, related citations] [Full Text]

  128. Stambolic, V., Ruel, L., Woodgett, J. R. Lithium inhibits glycogen synthase kinase-3 activity and mimics Wingless signalling in intact cells. Curr. Biol. 6: 1664-1668, 1996. Note: Erratum: Curr. Biol. 7: 196 only, 1997. [PubMed: 8994831, related citations] [Full Text]

  129. Stamer, K., Vogel, R., Thies, E., Mandelkow, E., Mandelkow, E.-M. Tau blocks traffic of organelles, neurofilaments, and APP vesicles in neurons and enhances oxidative stress. J. Cell Biol. 156: 1051-1063, 2002. [PubMed: 11901170, images, related citations] [Full Text]

  130. Stefansson, H., Helgason, A., Thorleifsson, G., Steinthorsdottir, V., Masson, G., Barnard, J., Baker, A., Jonasdottir, A., Ingason, A., Gudnadottir, V. G., Desnica, N., Hicks, A., and 15 others. A common inversion under selection in Europeans. Nature Genet. 37: 129-137, 2005. [PubMed: 15654335, related citations] [Full Text]

  131. Taniguchi, T., Doe, N., Matsuyama, S., Kitamura, Y., Mori, H., Saito, N., Tanaka, C. Transgenic mice expressing mutant (N279K) human tau show mutation dependent cognitive defects without neurofibrillary tangle formation. FEBS Lett. 579: 5704-5712, 2005. [PubMed: 16219306, related citations] [Full Text]

  132. Tatebayashi, Y., Miyasaka, T., Chui, D.-H., Akagi, T., Mishima, K., Iwasaki, K., Fujiwara, M., Tanemura, K., Murayama, M., Ishiguro, K., Planel, E., Sato, S., Hashikawa, T., Takashima, A. Tau filament formation and associative memory deficit in aged mice expressing mutant (R406W) human tau. Proc. Nat. Acad. Sci. 99: 13896-13901, 2002. [PubMed: 12368474, images, related citations] [Full Text]

  133. Taylor, L. M., McMillan, P. J., Liachko, N. F., Strovas, T. J., Ghetti, B., Bird, T. D., Keene, C. D., Kraemer, B. C. Pathological phosphorylation of tau and TDP-43 by TTBK1 and TTBK2 drives neurodegeneration. Molec. Neurodegener. 13: 7, 2018. [PubMed: 29409526, images, related citations] [Full Text]

  134. Tesseur, I., van Dorpe, J., Bruynseels, K., Bronfman, F., Sciot, R., van Lommel, A., van Leuven, F. Prominent axonopathy and disruption of axonal transport in transgenic mice expressing human apolipoprotein E4 in neurons of brain and spinal cord. Am. J. Path. 157: 1495-1510, 2000. [PubMed: 11073810, images, related citations] [Full Text]

  135. Tesseur, I., van Dorpe, J., Spittaels, K., van den Haute, C., Moechars, D., van Leuven, F. Expression of human apolipoprotein E4 in neurons causes hyperphosphorylation of protein tau in the brains of transgenic mice. Am. J. Path. 156: 951-964, 2000. [PubMed: 10702411, images, related citations] [Full Text]

  136. Tobin, J. E., Latourelle, J. C., Lew, M. F., Klein, C., Suchowersky, O., Shill, H. A., Golbe, L. I., Mark, M. H., Growdon, J. H., Wooten, G. F., Racette, B. A., Perlmutter, J. S., and 33 others. Haplotypes and gene expression implicate the MAPT region for Parkinson disease: the GenePD Study. Neurology 71: 28-34, 2008. [PubMed: 18509094, images, related citations] [Full Text]

  137. Tsuboi, Y., Baker, M., Hutton, M. L., Uitti, R. J., Rascol, O., Delisle, M.-B., Soulages, X., Murrell, J. R., Ghetti, B., Yasuda, M., Komure, O., Kuno, S., Arima, K., Sunohara, N., Kobayashi, T., Mizuno, Y., Wszolek, Z. K. Clinical and genetic studies of families with the tau N279K mutation (FTDP-17). Neurology 59: 1791-1793, 2002. [PubMed: 12473774, related citations] [Full Text]

  138. van Swieten, J. C., Stevens, M., Rosso, S. M., Rizzu, P., Joosse, M., de Koning, I., Kamphorst, W., Ravid, R., Spillantini, M. G., Niermeijer, Heutink, P. Phenotypic variation in hereditary frontotemporal dementia with tau mutations. Ann. Neurol. 46: 617-626, 1999. [PubMed: 10514099, related citations] [Full Text]

  139. Varani, L., Hasegawa, M., Spillantini, M. G., Smith, M. J., Murrell, J. R., Ghetti, B., Klug, A., Goedert, M., Varani, G. Structure of tau exon 10 splicing regulatory element RNA and destabilization by mutations of frontotemporal dementia and parkinsonism linked to chromosome 17. Proc. Nat. Acad. Sci. 96: 8229-8234, 1999. [PubMed: 10393977, images, related citations] [Full Text]

  140. Verpillat, P., Camuzat, A., Hannequin, D., Thomas-Anterion, C., Puel, M., Belliard, S., Dubois, B., Didic, M., Michel, B.-F., Lacomblez, L., Moreaud, O., Sellal, F., Golfier, V., Campion, D., Clerget-Darpoux, F., Brice, A. Association between the extended tau haplotype and frontotemporal dementia. Arch. Neurol. 59: 935-939, 2002. [PubMed: 12056929, related citations] [Full Text]

  141. Volz, A., Boyle, J. M., Cann, H. M., Cottingham, R. W., Orr, H. T., Ziegler, A. Report of the second international workshop in human chromosome 6. Genomics 21: 464-472, 1994. [PubMed: 8088851, related citations] [Full Text]

  142. Werber, E., Klein, C., Grunfeld, J., Rabey, J. M. Phenotypic presentation of frontotemporal dementia with parkinsonism-chromosome 17 type P301S in a patient of Jewish-Algerian origin. Mov. Disord. 18: 595-598, 2003. [PubMed: 12722177, related citations] [Full Text]

  143. Whitwell, J. L., Jack, C. R., Jr., Boeve, B. F., Senjem, M. L., Baker, M., Ivnik, R. J., Knopman, D. S., Wszolek, Z. K., Petersen, R. C., Rademakers, R., Josephs, K. A. Atrophy patterns in IVS10+16, IVS10+3, N279K, S305N, P301L, and V337M MAPT mutations. Neurology 73: 1058-1065, 2009. [PubMed: 19786698, images, related citations] [Full Text]

  144. Wilhelmsen, K. C., Lynch, T., Pavlou, E., Higgins, M., Hygaard, T. G. Localization of disinhibition-dementia-parkinsonism-amyotrophy complex to 17q21-22. Am. J. Hum. Genet. 55: 1159-1165, 1994. [PubMed: 7977375, related citations]

  145. Wszolek, Z. K., Pfeiffer, R. F., Bhatt, M. H., Schelper, R. L., Cordes, M., Snow, B. J., Rodnitzky, R. L., Wolters, E. C., Arwert, F., Calne, D. B. Rapidly progressive autosomal dominant parkinsonism and dementia with pallido-ponto-nigral degeneration. Ann. Neurol. 32: 312-320, 1992. [PubMed: 1416801, related citations] [Full Text]

  146. Wszolek, Z. K., Uitti, R. J., Hutton, M. A mutation in the microtubule-associated protein tau in pallido-nigro-luysian degeneration. (Letter) Neurology 54: 2028-2030, 2000. [PubMed: 10822460, related citations] [Full Text]

  147. Xu, J., Sato, S., Okuyama, S., Swan, R. J., Jacobsen, M. T., Strunk, E., Ikezu, T. Tau-tubulin kinase 1 enhances prefibrillar tau aggregation and motor neuron degeneration in P301L FTDP-17 tau-mutant mice. FASEB J. 24: 2904-2915, 2010. [PubMed: 20354135, related citations] [Full Text]

  148. Yasuda, M., Kawamata, T., Komure, O., Kuno, S., D'Souza, I., Poorkaj, P., Kawai, J., Tanimukai, S., Yamamoto, Y., Hasegawa, H., Sasahara, M., Hazama, F., Schellenberg, G. D., Tanaka, C. A mutation in the microtubule-associated protein tau in pallido-nigro-luysian degeneration. Neurology 53: 864-868, 1999. [PubMed: 10489057, related citations] [Full Text]

  149. Yasuda, M., Nakamura, Y., Kawamata, T., Kaneyuki, H., Maeda, K., Komure, O. Phenotypic heterogeneity within a new family with the MAPT P301S mutation. Ann. Neurol. 58: 920-928, 2005. [PubMed: 16240366, related citations] [Full Text]

  150. Yasuda, M., Yokoyama, K., Nakayasu, T., Nishimura, Y., Matsui, M., Yokoyama, T., Miyoshi, K., Tanaka, C. A Japanese patient with frontotemporal dementia and parkinsonism by a tau P301S mutation. Neurology 55: 1224-1227, 2000. [PubMed: 11071507, related citations] [Full Text]

  151. Yoshida, H., Crowther, R. A., Goedert, M. Functional effects of tau gene mutations deltaN296 and N296H. J. Neurochem. 80: 548-551, 2002. [PubMed: 11906000, related citations] [Full Text]

  152. Yoshiyama, Y., Higuchi, M., Zhang, B., Huang, S.-M., Iwata, N., Saido, T. C., Maeda, J., Suhara, T., Trojanowski, J. Q., Lee, V. M.-Y. Synapse loss and microglial activation precede tangles in a P301S tauopathy mouse model. Neuron 53: 337-351, 2007. Note: Erratum: Neuron 54: 343-344, 2007. [PubMed: 17270732, related citations] [Full Text]

  153. Zabetian, C. P., Hutter, C. M., Factor, S. A., Nutt, J. G., Higgins, D. S., Griffith, A., Roberts, J. W., Leis, B. C., Kay, D. M., Yearout, D., Montimurro, J. S., Edwards, K. L., Samii, A., Payami, H. Association analysis of MAPT H1 haplotype and subhaplotypes in Parkinson's disease. Ann. Neurol. 62: 137-144, 2007. [PubMed: 17514749, related citations] [Full Text]

  154. Zarranz, J. J., Ferrer, I., Lezcano, E., Forcadas, M. I., Eizaguirre, B., Atares, B., Puig, B., Gomez-Esteban, J. C., Fernandez-Maiztegui, C., Rouco, I., Perez-Concha, T., Fernandez, M., Rodriguez, O., Rodriguez-Martinez, A. B., Martinez de Pancorbo, M., Pastor, P., Perez-Tur, J. A novel mutation (K317M) in the MAPT gene causes FTDP and motor neuron disease. Neurology 64: 1578-1585, 2005. [PubMed: 15883319, related citations] [Full Text]

  155. Zhang, W., Tarutani, A., Newell, K. L., Murzin, A. G., Matsubara, T., Falcon, B., Vidal, R., Garringer, H. J., Shi, Y., Ikeuchi, T., Murayama, S., Ghetti, B., Hasegawa, M., Goedert, M., Scheres, S. H. W. Novel tau filament fold in corticobasal degeneration. Nature 580: 283-287, 2020. [PubMed: 32050258, images, related citations] [Full Text]

  156. Zhukareva, V., Mann, D., Pickering-Brown, S., Uryu, K., Shuck, T., Shah, K., Grossman, M., Miller, B. L., Hulette, C. M., Feinstein, S. C., Trojanowski, J. Q., Lee, V. M.-Y. Sporadic Pick's disease: a tauopathy characterized by a spectrum of pathological tau isoforms in gray and white matter. Ann. Neurol. 51: 730-739, 2002. [PubMed: 12112079, related citations] [Full Text]

  157. Zody, M. C., Jiang, Z., Fung, H.-C., Antonacci, F., Hillier, L. W., Cardone, M. F., Graves, T. A., Kidd, J. M., Cheng, Z., Abouelleil, A., Chen, L., Wallis, J., Glasscock, J., Wilson, R. J., Reily, A. D., Duckworth, J., Ventura, M., Hardy, J., Warren, W. C., Eichler, E. E. Evolutionary toggling of the MAPT 17q21.31 inversion region. Nature Genet. 40: 1076-1083, 2008. [PubMed: 19165922, images, related citations] [Full Text]


Bao Lige - updated : 11/09/2023
Bao Lige - updated : 06/30/2021
Bao Lige - updated : 05/05/2021
Ada Hamosh - updated : 08/14/2020
Ada Hamosh - updated : 08/11/2020
Ada Hamosh - updated : 01/06/2020
Ada Hamosh - updated : 09/12/2019
Ada Hamosh - updated : 11/19/2018
George E. Tiller - updated : 09/13/2017
Ada Hamosh - updated : 10/6/2015
George E. Tiller - updated : 9/4/2013
George E. Tiller - updated : 8/14/2013
Ada Hamosh - updated : 7/19/2012
Cassandra L. Kniffin - updated : 3/6/2012
Cassandra L. Kniffin - updated : 11/14/2011
Cassandra L. Kniffin - updated : 3/23/2011
Patricia A. Hartz - updated : 2/3/2011
Ada Hamosh - updated : 5/26/2010
Cassandra L. Kniffin - updated : 3/24/2010
George E. Tiller - updated : 2/22/2010
Cassandra L. Kniffin - updated : 1/4/2010
Cassandra L. Kniffin - updated : 12/14/2009
George E. Tiller - updated : 10/23/2009
Cassandra L. Kniffin - updated : 3/16/2009
George E. Tiller - updated : 1/12/2009
Ada Hamosh - updated : 10/30/2008
Cassandra L. Kniffin - updated : 10/17/2008
Cassandra L. Kniffin - updated : 4/18/2008
Ada Hamosh - updated : 4/4/2008
Cassandra L. Kniffin - updated : 1/7/2008
Cassandra L. Kniffin - updated : 10/11/2007
Ada Hamosh - updated : 5/30/2007
Victor A. McKusick - updated : 3/29/2007
Cassandra L. Kniffin - updated : 3/15/2007
George E. Tiller - updated : 10/5/2006
George E. Tiller - updated : 9/7/2006
Cassandra L. Kniffin - updated : 4/20/2006
Cassandra L. Kniffin - updated : 3/13/2006
Cassandra L. Kniffin - updated : 3/6/2006
Cassandra L. Kniffin -updated : 11/3/2005
Anne M. Stumpf - updated : 9/27/2005
Cassandra L. Kniffin - updated : 9/1/2005
Ada Hamosh - updated : 8/15/2005
Cassandra L. Kniffin - updated : 7/19/2005
Cassandra L. Kniffin - updated : 1/20/2005
Victor A. McKusick - updated : 12/9/2004
Cassandra L. Kniffin - updated : 10/27/2004
Victor A. McKusick - updated : 9/21/2004
Victor A. McKusick - updated : 9/9/2004
Cassandra L. Kniffin - reorganized : 6/10/2004
Cassandra L. Kniffin - updated : 1/7/2004
Victor A. McKusick - updated : 11/19/2003
Victor A. McKusick - updated : 9/8/2003
Cassandra L. Kniffin - updated : 4/29/2003
Cassandra L. Kniffin - updated : 3/6/2003
Cassandra L. Kniffin - updated : 2/11/2003
Cassandra L. Kniffin - updated : 1/21/2003
Victor A. McKusick - updated : 1/8/2003
Cassandra L. Kniffin - updated : 12/6/2002
Cassandra L. Kniffin - updated : 11/26/2002
Victor A. McKusick - updated : 11/22/2002
Dawn Watkins-Chow - updated : 11/5/2002
Cassandra L. Kniffin - updated : 10/2/2002
Dawn Watkins-Chow - updated : 8/22/2002
Cassandra L. Kniffin - updated : 7/29/2002
Victor A. McKusick - updated : 6/17/2002
Victor A. McKusick - updated : 6/6/2002
Cassandra L. Kniffin - updated : 6/5/2002
Victor A. McKusick - updated : 6/4/2002
Victor A. McKusick - updated : 12/4/2001
Dawn Watkins-Chow - updated : 11/25/2001
Ada Hamosh - updated : 9/12/2001
Victor A. McKusick - updated : 9/28/2000
Victor A. McKusick - updated : 7/31/2000
George E. Tiller - updated : 5/2/2000
Victor A. McKusick - updated : 3/24/2000
Ada Hamosh - updated : 2/22/2000
Victor A. McKusick - updated : 8/10/1999
Ada Hamosh - updated : 6/24/1999
Victor A. McKusick - updated : 4/7/1999
Victor A. McKusick - updated : 2/26/1999
Victor A. McKusick - updated : 2/18/1999
Victor A. McKusick - updated : 12/3/1998
Victor A. McKusick - updated : 11/18/1998
Victor A. McKusick - updated : 11/2/1998
Victor A. McKusick - updated : 8/20/1998
Victor A. McKusick - updated : 6/29/1998
Moyra Smith - updated : 7/1/1996
Andre K. Cheng - edited : 4/23/1996
Orest Hurko - updated : 4/4/1996
Creation Date:
Victor A. McKusick : 4/15/1987
mgross : 11/09/2023
mgross : 06/30/2021
carol : 05/06/2021
mgross : 05/05/2021
alopez : 08/14/2020
alopez : 08/11/2020
carol : 08/11/2020
carol : 01/07/2020
alopez : 01/06/2020
alopez : 09/12/2019
alopez : 11/19/2018
alopez : 09/13/2017
carol : 08/18/2016
joanna : 03/24/2016
carol : 1/11/2016
alopez : 10/6/2015
mgross : 9/17/2015
carol : 4/13/2015
alopez : 1/26/2015
alopez : 9/10/2013
tpirozzi : 9/4/2013
tpirozzi : 8/16/2013
tpirozzi : 8/16/2013
tpirozzi : 8/15/2013
tpirozzi : 8/15/2013
tpirozzi : 8/14/2013
terry : 4/4/2013
terry : 3/14/2013
terry : 11/29/2012
alopez : 7/23/2012
terry : 7/19/2012
carol : 6/8/2012
carol : 5/16/2012
carol : 3/9/2012
ckniffin : 3/6/2012
carol : 11/16/2011
terry : 11/16/2011
ckniffin : 11/14/2011
carol : 4/18/2011
carol : 4/14/2011
wwang : 4/12/2011
ckniffin : 3/23/2011
mgross : 2/9/2011
terry : 2/3/2011
wwang : 1/7/2011
ckniffin : 12/10/2010
alopez : 12/8/2010
ckniffin : 11/17/2010
alopez : 6/1/2010
terry : 5/26/2010
carol : 3/24/2010
ckniffin : 3/24/2010
wwang : 3/2/2010
terry : 2/22/2010
wwang : 1/20/2010
ckniffin : 1/4/2010
wwang : 12/28/2009
carol : 12/23/2009
ckniffin : 12/15/2009
ckniffin : 12/14/2009
wwang : 11/3/2009
terry : 10/23/2009
wwang : 8/24/2009
terry : 8/12/2009
terry : 6/3/2009
alopez : 4/15/2009
wwang : 3/26/2009
ckniffin : 3/16/2009
wwang : 1/12/2009
alopez : 10/30/2008
wwang : 10/27/2008
ckniffin : 10/17/2008
wwang : 4/23/2008
ckniffin : 4/18/2008
alopez : 4/11/2008
terry : 4/4/2008
carol : 2/29/2008
wwang : 1/23/2008
ckniffin : 1/7/2008
wwang : 10/19/2007
ckniffin : 10/11/2007
alopez : 5/30/2007
terry : 5/30/2007
terry : 5/9/2007
carol : 3/29/2007
carol : 3/29/2007
ckniffin : 3/15/2007
alopez : 10/5/2006
alopez : 9/7/2006
carol : 6/21/2006
ckniffin : 6/12/2006
wwang : 4/25/2006
ckniffin : 4/20/2006
alopez : 4/12/2006
wwang : 3/20/2006
ckniffin : 3/13/2006
wwang : 3/13/2006
ckniffin : 3/6/2006
alopez : 2/3/2006
wwang : 11/11/2005
ckniffin : 11/3/2005
alopez : 9/27/2005
wwang : 9/6/2005
ckniffin : 9/1/2005
carol : 8/16/2005
terry : 8/15/2005
wwang : 7/26/2005
ckniffin : 7/19/2005
terry : 7/11/2005
terry : 2/22/2005
tkritzer : 1/25/2005
ckniffin : 1/20/2005
tkritzer : 1/6/2005
terry : 12/9/2004
terry : 11/2/2004
tkritzer : 11/1/2004
ckniffin : 10/27/2004
tkritzer : 9/23/2004
terry : 9/21/2004
tkritzer : 9/10/2004
terry : 9/9/2004
carol : 6/10/2004
ckniffin : 6/9/2004
ckniffin : 6/8/2004
mgross : 3/17/2004
carol : 1/29/2004
ckniffin : 1/7/2004
tkritzer : 12/19/2003
tkritzer : 12/18/2003
tkritzer : 11/24/2003
terry : 11/19/2003
cwells : 9/10/2003
terry : 9/8/2003
tkritzer : 5/21/2003
tkritzer : 4/29/2003
ckniffin : 4/29/2003
carol : 3/17/2003
tkritzer : 3/14/2003
ckniffin : 3/6/2003
carol : 2/25/2003
carol : 2/25/2003
ckniffin : 2/11/2003
carol : 1/31/2003
tkritzer : 1/28/2003
ckniffin : 1/21/2003
cwells : 1/9/2003
tkritzer : 1/8/2003
carol : 12/6/2002
ckniffin : 12/6/2002
cwells : 12/3/2002
ckniffin : 11/26/2002
cwells : 11/22/2002
terry : 11/20/2002
carol : 11/7/2002
tkritzer : 11/5/2002
tkritzer : 11/5/2002
carol : 10/31/2002
tkritzer : 10/25/2002
carol : 10/21/2002
ckniffin : 10/4/2002
ckniffin : 10/4/2002
ckniffin : 10/2/2002
tkritzer : 8/22/2002
carol : 8/7/2002
ckniffin : 8/7/2002
ckniffin : 7/29/2002
mgross : 6/25/2002
terry : 6/21/2002
terry : 6/17/2002
alopez : 6/12/2002
mgross : 6/10/2002
mgross : 6/10/2002
terry : 6/6/2002
carol : 6/5/2002
ckniffin : 6/5/2002
ckniffin : 6/5/2002
terry : 6/4/2002
carol : 12/10/2001
mcapotos : 12/4/2001
carol : 11/25/2001
alopez : 9/13/2001
alopez : 9/13/2001
terry : 9/12/2001
alopez : 10/30/2000
mcapotos : 10/17/2000
mcapotos : 10/16/2000
terry : 10/6/2000
terry : 9/28/2000
alopez : 8/31/2000
alopez : 7/31/2000
terry : 7/31/2000
alopez : 5/2/2000
mcapotos : 4/18/2000
mcapotos : 4/14/2000
terry : 3/24/2000
alopez : 2/22/2000
alopez : 8/23/1999
terry : 8/10/1999
alopez : 8/5/1999
alopez : 6/24/1999
alopez : 6/24/1999
carol : 4/7/1999
terry : 2/26/1999
carol : 2/22/1999
terry : 2/18/1999
alopez : 12/4/1998
alopez : 12/3/1998
terry : 11/18/1998
carol : 11/11/1998
carol : 11/10/1998
terry : 11/2/1998
terry : 8/20/1998
dholmes : 7/22/1998
carol : 6/29/1998
terry : 6/29/1998
mark : 9/19/1996
mark : 7/1/1996
terry : 7/1/1996
terry : 7/1/1996
mark : 4/23/1996
mark : 4/23/1996
terry : 4/15/1996
mark : 4/4/1996
carol : 9/22/1993
supermim : 3/16/1992
supermim : 3/20/1990
ddp : 10/27/1989
root : 6/27/1988
root : 6/20/1988

* 157140

MICROTUBULE-ASSOCIATED PROTEIN TAU; MAPT


Alternative titles; symbols

MTBT1


HGNC Approved Gene Symbol: MAPT

SNOMEDCT: 230270009;   ICD10CM: G31.0, G31.01;   ICD9CM: 331.1, 331.11;  


Cytogenetic location: 17q21.31     Genomic coordinates (GRCh38): 17:45,894,554-46,028,334 (from NCBI)


Gene-Phenotype Relationships

Location Phenotype Phenotype
MIM number
Inheritance Phenotype
mapping key
17q21.31 {Parkinson disease, susceptibility to} 168600 Autosomal dominant; Multifactorial 3
Dementia, frontotemporal, with or without parkinsonism 600274 Autosomal dominant 3
Pick disease 172700 Autosomal dominant 3
Supranuclear palsy, progressive 601104 Autosomal dominant 3
Supranuclear palsy, progressive atypical 260540 Autosomal recessive 3

TEXT

Cloning and Expression

The microtubule-associated proteins (MAPs) coassemble with tubulin (see 602529) into microtubules in vitro. Microtubule-associated protein tau appears to be enriched in axons. Neve et al. (1986) identified tau cDNA clones in a human fetal brain cDNA library. The clones recognized a 6-kb message that was expressed in human brain but not in other human tissues and exhibited a developmental shift in size.

By screening cDNA libraries prepared from the frontal cortex of an Alzheimer disease patient and from fetal human brain, Goedert et al. (1988) isolated the cDNA for a core protein of the paired helical filament of Alzheimer disease (AD; 104300). The partial amino acid sequence of this core protein was used to design synthetic oligonucleotide probes. The cDNA encodes a protein of 352 amino acids that contains a characteristic amino acid repeat in its carboxyl-terminal half. Because of extensive homology to the sequence of the mouse microtubule-associated protein tau, they stated that this protein must constitute the human equivalent of mouse tau. Tau protein mRNA was found in normal amounts in the frontal cortex from patients with Alzheimer disease.

Goedert et al. (1989) determined the sequences of 6 tau isoforms produced in adult human brain by alternative mRNA splicing. The proteins are composed of 352 to 441 amino acids. The isoforms differ from each other by the presence or absence of 29-amino acid or 58-amino acid inserts located in the N terminus and a 31-amino repeat located in the C terminus. Inclusion of the latter, which is encoded by exon 10 of the tau gene, gives rise to the 3 tau isoforms with 4 repeats each; the other 3 isoforms have 3 repeats each. Normal cerebral cortex contains similar levels of 3-repeat and 4-repeat tau isoforms. The repeats and some adjoining sequences constitute the microtubule-binding domains of tau (see also Hutton et al., 1998).

Holzer et al. (2004) determined that the longest form (441 amino acids) of the chimpanzee brain tau protein shares 100% amino acid identity with human tau. The nonconstitutively spliced exon 4A differed at 3 of 251 amino acid positions. Identities for gorilla and gibbon tau with human tau were 99.5% and 99.0%, respectively. Analysis of 8 polymorphic markers revealed that the nonhuman primates had a higher prevalence of the H2 haplotype, in contrast to humans, in whom the H1 haplotype is more common. Holzer et al. (2004) noted that chimpanzee brains show resistance to developing tau pathology and suggested that differences in intronic sequences of the tau gene may be responsible.


Gene Structure

Andreadis et al. (1992) found that the tau isoforms that predominate in human brain are encoded by 11 exons.

Conrad et al. (2002) identified an intronless gene, saitohin (STH; 607067), located in the intron between exons 9 and 10 of the tau gene.


Mapping

The MAPT gene was assigned to chromosome 17 by hybridization of a cDNA clone to flow-sorted and spot-blotted chromosomes (Neve et al., 1986) and to 17q21 by in situ hybridization (Donlon et al., 1987).

In the course of constructing a radiation hybrid map of the breast cancer (113705) region of 17q, Abel et al. (1993) concluded that the tau protein gene (symbolized MTBT1 by them) is proximal to GP3A (173470), located at 17q21.32, and distal to TOP2A (126430) and THRA1 (190120), located at 17q11.2. ERBB2 (164870), previously localized to 17q21-q22, was found to be proximal to all of these. Poorkaj et al. (2001) identified and sequenced a human PAC and a mouse BAC containing the entire MAPT and Mtapt genes, respectively. They found that the corticotropin-releasing hormone receptor gene (CRHR1; 122561) is the next gene 5-prime to MAPT. The gene located 3-prime to MAPT and encoded by the opposite DNA strand has a predicted sequence identical to the KIAA1267 cDNA (KANSL1; 612452) identified from adult human brain by Nagase et al. (1999).


Gene Function

Alonso et al. (1996) reported studies on the microtubule-associated protein tau in Alzheimer disease. They noted that in the brains of patients with Alzheimer disease the neuronal cytoskeleton is progressively disrupted and replaced by tangles of paired helical filaments (PHFs), and that PHFs are composed mainly of hyperphosphorylated forms of tau (called 'AD P-tau' by them). They demonstrated that in solution normal tau associated with the hyperphosphorylated AD P-tau to form large tangles of filaments. They also demonstrated that dephosphorylation with alkaline phosphatase abolished the ability of AD P-tau to aggregate in vitro.

Hypothesizing that restoring the function of phosphorylated tau might prevent or reverse PHF formation in Alzheimer disease, and with the knowledge that PIN1 (601052) specifically isomerizes phosphorylation of a serine or threonine that precedes proline and regulates the function of mitotic phosphoproteins, Lu et al. (1999) demonstrated that the WW domain of PIN1 binds to phosphorylated tau at threonine-231 (T231). The T231 residue is hyperphosphorylated in Alzheimer disease and is phosphorylated to a certain extent in the normal brain. Using a pull-down assay, Lu et al. (1999) demonstrated that PIN1 binds to hyperphosphorylated tau from the brains of people with Alzheimer disease but not to tau from age-matched healthy brains. By immunoblotting, Lu et al. (1999) detected endogenous PIN1 in the PHFs of diseased brains, and using immunohistochemistry, they found that recombinant PIN1 binds to pathologic tau. Using immunohistochemistry, Lu et al. (1999) localized PIN1 to the nucleus in healthy brains. In the brains of people with Alzheimer disease, PIN1 staining was associated with pathologic tau in neuronal cells. Lu et al. (1999) also demonstrated that phosphorylated tau could neither bind microtubules nor promote microtubule assembly. However, PIN1 was able to restore the ability of phosphorylated tau to bind microtubules and promoted microtubule assembly in vitro. The level of soluble PIN1 in the brains of Alzheimer patients was greatly reduced compared to that in age-matched control brains. The authors concluded with the hypothesis that since depletion of PIN1 induces mitotic arrest and apoptotic cell death, sequestration of PIN1 into PHFs may contribute to neuronal death.

Schneider et al. (1999) presented evidence that challenged the common hypothesis that hyperphosphorylated tau promotes aggregation into PHFs. Using an in vitro system, they showed that treatment with several different kinases phosphorylated tau and caused detachment from microtubules, yet at the same time protected against the formation of PHFs. The authors suggested the findings are due most likely to the complex interaction of tau's ability to adopt many conformations and the presence of phosphorylation sites with differing affinities and locations.

Senile plaques and neurofibrillary tangles, the 2 hallmark lesions of Alzheimer disease, are the result of the pathologic deposition of proteins normally present throughout the brain. Senile plaques are extracellular deposits of fibrillar beta-amyloid peptide; neurofibrillary tangles represent intracellular bundles of self-assembled hyperphosphorylated tau proteins. These 2 lesions are often present in the same brain areas, and Rapoport et al. (2002) presented work that suggested a mechanistic link between them. They analyzed whether tau plays a key role in fibrillar beta-amyloid-induced neurite degeneration in central nervous system neurons. Cultured hippocampal neurons obtained from wildtype, tau knockout, and human tau transgenic mice were treated with fibrillar amyloid-beta. Morphologic analysis indicated that neurons expressing either mouse or human tau proteins degenerated in the presence of amyloid-beta. On the other hand, tau-depleted neurons showed no signs of degeneration in the presence of amyloid-beta.

From detailed analysis of pathologic load and spatiotemporal distribution of beta-amyloid deposits and tau pathology in sporadic Alzheimer disease, Delacourte et al. (2002) concluded that there is a synergetic effect of amyloid aggregation in the propagation of tau pathology.

The association of intronic mutations in the MAPT gene (e.g., 157140.0004) in frontotemporal dementia with parkinsonism (600274) highlights the involvement of aberrant pre-mRNA splicing in the pathogenesis of neurodegenerative disorders. To establish a model system for studying the role of pre-mRNA splicing in neurodegenerative diseases, Jiang et al. (2000) constructed a MAPT minigene that reproduced alternative splicing in both cultured cells and in vitro biochemical assays. They demonstrated that mutations in a nonconserved intronic region of the human MAPT gene led to increased splicing between exons 10 and 11. Systematic biochemical analyses indicated the importance of U1 snRNP (180740) and, to a lesser extent, U6 snRNP (180692) in differentially recognizing wildtype versus intron-mutant MAPT pre-mRNAs.

Goode et al. (2000) analyzed the structure and function of the 3-repeat (3R) and 4-repeat (4R) tau isoforms. Rather than having linear, sequentially arranged tubulin-binding domains, the isoforms have specific core microtubule-binding domains that lead to complex intramolecular folding interactions. Flanking regions were also found to contribute to the binding activity in the 3-repeat isoform, but less so in the 4-repeat isoform. Goode et al. (2000) suggested that the 2 isoforms form distinct structures that likely have different functional capabilities. Panda et al. (2003) noted that the abnormally high ratio of 4R to 3R tau in the MAPT gene might lead to neuronal cell death by altering normal tau functions in adult neurons. They tested whether 3R and 4R tau might differentially modulate the dynamic instability of microtubules in vitro using video microscopy. Although both isoforms promoted microtubule polymerization and decreased the tubulin critical subunit concentration to approximately similar extents, 4R tau stabilized microtubules significantly more strongly than 3R tau. Panda et al. (2003) suggested a 'dosage effect' or haploinsufficiency model in which both tau alleles must be active and properly regulated to produce appropriate amounts of each tau isoform to maintain microtubule dynamics within a tolerable window of activity.

Stamer et al. (2002) showed that elevated levels of tau inhibit intracellular transport in neurons, particularly the plus-end-directed transport by kinesin motors from the center of the cell body to the neuronal processes. This inhibition is significant because critical organelles, such as peroxisomes, mitochondria, and transport vesicles carrying supplies for the growth cone, are unable to penetrate the neurites, leading to stunted growth, increased susceptibility to oxidative stress, and likely pathologic aggregation of proteins such as amyloid precursor protein (APP; 104760). Stamer et al. (2002) concluded that the tau:tubulin ratio is normally low, and that increased levels of tau become detrimental to the cell.

Giasson et al. (2003) showed that alpha-synuclein (SNCA; 163890) induces fibrillization of tau, and that coincubation of alpha-synuclein and tau synergistically promotes fibrillization of both proteins in vitro. In vivo studies of mice with an alpha-synuclein mutation or a tau mutation showed filamentous inclusions of both proteins, which are abundant neuronal proteins that normally adopt an unfolded conformation but polymerize into amyloid fibrils in disease. The findings suggested an interaction between alpha-synuclein and tau that drives the formation of pathologic inclusions in human neurodegenerative diseases.

Rizzu et al. (2004) presented evidence suggesting that DJ1 (602533) colocalizes within a subset of pathologic tau inclusions in tauopathies, and that the solubility of DJ1 is altered in association with its aggregation within these inclusions.

Petrucelli et al. (2004) reported that CHIP (607207), a ubiquitin ligase that interacts directly with Hsp70/90 (140550/140571), induced ubiquitination and increased aggregation of tau. Tau lesions in human postmortem tissue were immunopositive for CHIP. Conversely, induction of Hsp70 through treatment with either geldanamycin or heat shock factor-1 (HSF1; 140580) led to a decrease in tau steady-state levels and a selective reduction in detergent insoluble tau. Furthermore, 30-month-old mice overexpressing inducible Hsp70 showed a significant reduction in tau levels. The authors concluded that the Hsp70/CHIP chaperone system may play an important role in the regulation of tau turnover and the selective elimination of abnormal tau species.

Liu et al. (2004) investigated the mechanisms leading to abnormal hyperphosphorylation of tau in pathologic states. They demonstrated that human brain tau was modified by O-GlcNAcylation, a type of protein O-glycosylation by which the monosaccharide beta-N-acetylglucosamine (GlcNAc) attaches to serine/threonine residues via an O-linked glycosidic bond. This glycosylation regulated tau phosphorylation in a site-specific manner both in vitro and in vivo. At most of the phosphorylation sites, the process negatively regulated tau phosphorylation. In an animal model of starved mice, low glucose uptake/metabolism that mimicked those observed in Alzheimer disease brain produced a decrease in O-GlcNAcylation and consequent hyperphosphorylation of tau at the majority of the phosphorylation sites. The O-GlcNAcylation level in Alzheimer disease brain extracts was decreased as compared to that in controls.

Using Western blotting, immunoprecipitation assays, and surface plasmon resonance analysis, Guo et al. (2006) showed that beta-amyloid-40 and -42 formed stable complexes with soluble tau and that prior phosphorylation of tau inhibited complex formation. Immunostaining of brain extracts from patients with AD and controls showed that phosphorylated tau and beta-amyloid were present within the same neuron. Guo et al. (2006) postulated that an initial step in AD pathogenesis may be the intracellular binding of soluble beta-amyloid to soluble nonphosphorylated tau.

In studies of rodent and human cells, Li et al. (2007) found that overexpression of hyperphosphorylated tau antagonized apoptosis of neuronal cells by stabilizing beta-catenin (CTNNB1; 116806). The findings explained why NFT-bearing neurons survive proapoptotic insults and instead die chronically of degeneration.

Roberson et al. (2007) found that reducing endogenous tau levels prevented behavioral deficits in transgenic mice expressing human APP, without altering their high A-beta levels. Tau reduction also protected both transgenic and nontransgenic mice against excitotoxicity. Roberson et al. (2007) concluded that thus, tau reduction can block A-beta- and excitotoxin-induced neuronal dysfunction and may represent an effective strategy for treating Alzheimer disease and related conditions.

To determine the effects of tau on dynein (600112) and kinesin (602809) motility, Dixit et al. (2008) conducted single-molecule studies of motor proteins moving along tau-decorated microtubules. Dynein tended to reverse direction, whereas kinesin tended to detach at patches of bound tau. Kinesin was inhibited at about a tenth of the tau concentration that inhibited dynein, and the microtubule-binding domain of tau was sufficient to inhibit motor activity. The differential modulation of dynein and kinesin motility suggested that microtubule-associated proteins (MAPs) can spatially regulate the balance of microtubule-dependent axonal transport.

De Calignon et al. (2010) used in vivo multiphoton imaging to observe tangles and activation of executioner caspases in living tau transgenic mice (Tg4510 strain) created by SantaCruz et al. (2005), and found that caspase activation occurs prior to tangle formation and precedes tangle formation by hours to days. New tangles form within a day. After a new tangle forms, the neuron remains alive and caspase activity seems to be suppressed. Similarly, introduction of wildtype 4-repeat tau (tau-4R) into wildtype animals triggered caspase activation, tau truncation, and tau aggregation. Adeno-associated virus-mediated expression of a construct mimicking caspase-cleaved tau into wildtype mice led to the appearance of intracellular aggregates, tangle-related conformational- and phospho-epitopes, and the recruitment of full-length endogenous tau to the aggregates. On the basis of these data, de Calignon et al. (2010) proposed a new model in which caspase activation cleaves tau to initiate tangle formation, then truncated tau recruits normal tau to misfold and form tangles. Because tangle-bearing neurons are long-lived, de Calignon et al. (2010) suggested that tangles are 'off pathway' to acute neuronal death. De Calignon et al. (2010) suggested that soluble tau species, rather than fibrillar tau, may be the critical toxic moiety underlying neurodegeneration.

Amino-terminally truncated, pyroglutamylated (pE) forms of amyloid-beta (104760) are strongly associated with Alzheimer disease, are more toxic than amyloid-beta(1-42) and amyloid-beta(1-40), and have been proposed as initiators of Alzheimer disease pathogenesis. Nussbaum et al. (2012) reported a mechanism by which pE-amyloid-beta may trigger Alzheimer disease. Amyloid-beta-3(pE)-42 co-oligomerizes with excess amyloid-beta(1-42) to form metastable low-n oligomers (LNOs) that are structurally distinct and far more cytotoxic to cultured neurons than comparable LNOs made from amyloid-beta(1-42) alone. Tau is required for cytotoxicity, and LNOs comprising 5% amyloid-beta-3(pE)-42 plus 95% amyloid-beta(1-42) (5% pE-amyloid-beta) seed new cytotoxic LNOs through multiple serial dilutions into amyloid-beta(1-42) monomers in the absence of additional amyloid-beta-3(pE)-42. LNOs isolated from human Alzheimer disease brain contained amyloid-beta-3(pE)-42, and enhanced amyloid-beta-3(pE)-42 formation in mice triggered neuron loss and gliosis at 3 months, but not in a tau-null background. Nussbaum et al. (2012) concluded that amyloid-beta-3(pE)-42 confers tau-dependent neuronal death and causes template-induced misfolding of amyloid-beta(1-42) into structurally distinct LNOs that propagate by a prion-like mechanism. Nussbaum et al. (2012) concluded that their results raised the possibility that amyloid-beta-3(pE)-42 acts similarly at a primary step in Alzheimer disease pathogenesis.

Using in situ hybridization and immunohistochemical analyses, Lund et al. (2013) showed that TTBK1 (619415) was a neuron-specific kinase in human brain, with primary expression in the somatodendritic compartment. TTBK1 phosphorylated tau at ser422 and colocalized with phosphorylated ser422 pretangle neurons, but not with late-stage neurofibrillary tangles.

Kondo et al. (2015) investigated how traumatic brain injury (TBI), an environmental risk factor for Alzheimer disease, leads to tauopathy. Kondo et al. (2015) found robust cis phosphorylated tau protein (P-tau) pathology after TBI in humans and mice. After TBI in mice and stress in vitro, neurons acutely produced cis P-tau, which disrupted axonal microtubule networks and mitochondrial transport, spread to other neurons, and led to apoptosis. This process, which they termed 'cistauosis,' appears long before other tauopathy. Treating TBI mice with cis antibody blocked cistauosis, prevented tauopathy development and spread, and restored many TBI-related structural and functional sequelae. Thus, Kondo et al. (2015) concluded that cis P-tau is a major early driver of disease after TBI and leads to tauopathy in traumatic encephalopathy and Alzheimer disease. The authors suggested that the cis antibody may be further developed to detect and treat TBI and prevent progressive neurodegeneration after injury.

Faraco et al. (2019) reported that dietary salt induced hyperphosphorylation of tau followed by cognitive dysfunction in mice, and that these effects were prevented by restoring endothelial nitric oxide production. The nitric oxide deficiency reduced neuronal calpain (see 114220) nitrosylation and resulted in enzyme activation, which, in turn, led to tau phosphorylation by activating cyclin-dependent kinase-5 (CDK5; 123831). Salt-induced cognitive impairment was not observed in tau-null mice or in mice treated with anti-tau antibodies, despite persistent cerebral hypoperfusion and neurovascular dysfunction. Faraco et al. (2019) concluded that these findings identified a causal link between dietary salt, endothelial dysfunction, and tau pathology, independent of hemodynamic insufficiency. They further suggested that avoidance of excessive salt intake and maintenance of vascular health may help to stave off the vascular and neurodegenerative pathologies that underlie dementia in the elderly.

Rauch et al. (2020) showed that the low density lipoprotein receptor-related protein-1 (LRP1; 107770) controls the endocytosis of tau and its subsequent spread. Knockdown of LRP1 significantly reduced tau uptake in H4 neuroglioma cells and in induced pluripotent stem cell-derived neurons. The interaction between tau and LRP1 is mediated by lysine residues in the microtubule-binding repeat region of tau. Furthermore, downregulation of LRP1 in an in vivo mouse model of tau spread was found to effectively reduce the propagation of tau between neurons. Rauch et al. (2020) concluded that their results identified LRP1 as a key regulator of tau spread in the brain.


Biochemical Features

Cryoelectron Microscopy

Using cryoelectron microscopy, Falcon et al. (2018) determined the structures of tau filaments from patients with Pick disease (172700), a neurodegenerative disorder characterized by frontotemporal dementia. The filaments consist of residues lys254-phe378 of 3R tau, which are folded differently from the tau filaments in Alzheimer disease, establishing the existence of conformers of assembled tau. The observed tau fold in the filaments of patients with Pick disease explains the selective incorporation of 3R tau in Pick bodies, and the differences in phosphorylation relative to the tau filaments of Alzheimer disease. Falcon et al. (2018) concluded that their findings showed how tau can adopt distinct folds in the human brain in different diseases, an essential step for understanding the formation and propagation of molecular conformers.

Using cryoelectron microscopy, Falcon et al. (2019) determined the structure of tau filaments from the brains of 3 individuals with chronic traumatic encephalopathy (CTE) at resolutions down to 2.3 angstroms. Falcon et al. (2019) showed that filament structures were identical in the 3 cases but were distinct from those of Alzheimer disease and Pick disease, and from those formed in vitro. Similar to Alzheimer disease, all 6 brain tau isoforms assemble into filaments in CTE, and residues K274-R379 of 3-repeat tau and S305-R379 of 4-repeat tau form the ordered core of 2 identical C-shaped protofilaments. However, a different conformation of the beta-helix region creates a hydrophobic cavity that is absent in tau filaments from the brains of patients with Alzheimer disease. This cavity encloses an additional density that is not connected to tau, which suggests that the incorporation of cofactors may have a role in tau aggregation in CTE. Moreover, filaments in CTE have distinct protofilament interfaces to those of Alzheimer disease. Falcon et al. (2019) concluded that their structures provided a unifying neuropathologic criterion for CTE, and supported the hypothesis that the formation and propagation of distinct conformers of assembled tau underlie different neurodegenerative diseases.

Using cryoelectron microscopy, Zhang et al. (2020) analyzed the structures of tau filaments extracted from the brains of 3 individuals with corticobasal degeneration (CBD). These filaments were identical between cases, but distinct from those seen in Alzheimer disease (104300), Pick disease (172700), and chronic traumatic encephalopathy. The core of a CBD filament comprises residues lysine-274 to glutamate-380 of tau, spanning the last residue of the R1 repeat, the whole of the R2, R3, and R4 repeats, and 12 amino acids after R4. The core adopts a previously unseen 4-layered fold, which encloses a large nonproteinaceous density. This density is surrounded by the side chains of lysine residues 290 and 294 from R2 and lysine-370 from the sequence after R4.


Population Genetics

The MAPT gene has a polymorphic inversion resulting in 2 main haplotypes: the H1 haplotype, which comprises the more common human noninverted sequence, and the H2 haplotype, which comprises an inverted sequence and includes several other genes within its 900-kb extent (review by Donnelly et al., 2010). The promoter in H1 chromosomes is more efficient at driving transcription than the promoter sequence on the H2 haplotype (Kwok et al., 2004).

Studies by Baker et al. (1999), Skipper et al. (2004), and others identified extensive linkage disequilibrium in the 17q21.31 region and considerable divergence between the H1 and H2 lineages of MAPT. Using a refined physical map of chromosome 17q21.31, Stefansson et al. (2005) uncovered a 900-kb inversion polymorphism in this region accounting for these observations. Chromosomes with the inverted segment in different orientations represent the 2 distinct lineages H1 and H2 that have diverged for as much as 3 million years and show no evidence of having recombined. The H2 lineage is rare in Africans, almost absent in East Asians, but is found at a rate of approximately 20% in Europeans, in whom the haplotype structure is indicative of a history of positive selection. The frequency is about 6% in Africans and less than 1% in East Asians. Stefansson et al. (2005) showed that the H2 lineage is undergoing positive selection in the Icelandic population, such that carrier females have more children and have higher recombination rates than noncarriers.

Zody et al. (2008) used comparative sequencing approaches to investigate the evolutionary history of the European-enriched 17q21.31 MAPT inversion polymorphism. The authors presented a detailed BAC-based sequence assembly of the inverted human H2 haplotype and compared it to the sequence structure and genetic variation of the corresponding 1.5-Mb region for the noninverted H1 human haplotype and that of chimpanzee and orangutan. Zody et al. (2008) found that the inversion of the MAPT region is similarly polymorphic in other great ape species, and presented evidence that the inversions occurred independently in chimpanzees and humans. In humans, the inversion breakpoints correspond to core duplications with the LRRC37 gene family (see 616555). The analysis of Zody et al. (2008) favored the H2 configuration and sequence haplotype as the likely great ape and human ancestral state, with inversion recurrences during primate evolution. The authors further showed that the H2 architecture has evolved more extensive sequence homology, perhaps explaining its tendency to undergo microdeletion associated with mental retardation in European populations.

Donnelly et al. (2010) genotyped SNPs in the MAPT gene corresponding to the H1 and H2 haplotypes in 3,135 individuals from 66 populations worldwide. The H2 inversion was found at the highest frequencies in Southwest Asia and Southern Europe (approximately 30%). Elsewhere in Europe, the frequency of H2 varied from less than 5% in Finns to 28% in Orcadians. The H2 inversion haplotype occurs at low frequencies (1 to 10%) in Africa, Central Asia, East Asia, and the Americas, although the East Asian and Amerindian alleles may be due to recent gene flow from Europe. Molecular evolution analyses indicated that the H2 haplotype originally arose in Africa or Southwest Asia. The most recent common human ancestor was estimated to be from 13,600 to 108,400 years ago, which is much more recent than the 3 million year age estimated by Stefansson et al. (2005).


Molecular Genetics

Rademakers et al. (2004) reported that in the previous 5 years, research had identified 34 different pathogenic MAPT mutations in 101 families worldwide. They described the considerable differences in clinical and pathologic presentation of patients with MAPT mutations and summarized the effect of the different mutations on tau functioning. In addition, they discussed the role of tau as a genetic susceptibility factor, together with the genetic evidence for additional causal genes for tau-positive as well tau-negative dementia.

Frontotemporal Dementia with Parkinsonism

In 13 families with autosomal dominant frontotemporal dementia with parkinsonism linked to chromosome 17 (FTDP17; 600274), Hutton et al. (1998) identified mutations in the MAPT gene: 3 were missense mutations (157140.0001-157140.0003) and 3 were mutations in the 5-prime splice site of exon 10 (157140.0004-157140.0006). All of the splice site mutations destabilized a potential stem-loop structure likely involved in regulating the alternative splicing of exon 10 (Goedert et al., 1989), resulting in more frequent usage of the 5-prime splice site and an increased proportion of tau transcripts that included exon 10. The increase in exon 10+ mRNA was expected to increase the proportion of tau containing 4 microtubule-binding repeats, consistent with the neuropathology described in families with FTDP17. Most cases showed neuronal and/or glial inclusions that stained positively with antibodies raised against tau, although the tau pathology varied considerably in both its quantity and characteristics; the disorder is one of several termed 'tauopathies.'

Varani et al. (1999) determined the 3-dimensional structure of the tau exon 10 splicing regulatory element RNA by means of NMR spectroscopy. They showed that it does indeed form a stable, folded stem-loop structure whose thermodynamic stability is reduced by FTDP17 mutations and increased by compensatory mutations. By exon trapping, Varani et al. (1999) showed that the reduction in thermodynamic stability was correlated with increased splicing of exon 10.

Hong et al. (1998) indicated that more than 10 exonic and intronic mutations of the MAPT gene had been identified in about 20 FTDP17 families. Analyses of soluble and insoluble tau proteins from brains of FTDP17 patients indicated that different pathogenic mutations differentially altered distinct biochemical properties and stoichiometry of brain tau isoforms. Functional assays of recombinant tau proteins with different FTDP17 missense mutations implicated all but 1 of these mutations in disease pathogenesis by reducing the ability of tau to bind microtubules and promote microtubule assembly.

In a study of frontotemporal dementia in the Netherlands from January 1994 to June 1998, Rizzu et al. (1999) found 37 patients who had one or more first-degree relatives with dementia. A mutation in the MAPT gene was found in 17.8% of the group of patients with FTDP17 and in 43% of patients with FTDP17 who also had a positive family history of the disorder. Three distinct missense mutations, G272V (157140.0002), P301L (157140.0001), and R406W (157140.0003) accounted for 15.6% of the mutations. The missense mutations and a single amino acid deletion that was detected in 1 patient, strongly reduced the ability of tau to promote microtubule assembly. Rizzu et al. (1999) suggested that the MAPT mutations cause disturbances in the interactions of the tau protein with microtubules, resulting in hyperphosphorylation of tau protein, assembly into filaments, and subsequent cell death.

Verpillat et al. (2002) found that the tau H1/H1 genotype (see below) was significantly overrepresented in 100 patients with frontotemporal dementia compared to controls (odds ratio for H1/H1 = 1.95). In addition, there was a significant negative effect in carriers of both the H1/H1 genotype and the APOE2 allele (107741).

Goedert et al. (1998) reviewed the role of tau mutations in frontotemporal dementias. Heutink (2000) reviewed the role of tau protein in frontotemporal dementia and other neurodegenerative disorders. Hutton (2001) reviewed the known missense and splice site mutations in the tau gene that are associated with disease and described different mechanisms involved in pathogenesis, including disruption of the interaction between tau and tubulin, deposition of abnormal tau filaments, and the generation of abnormal ratios of tau isoforms.

Using purified recombinant proteins, Alonso et al. (2004) showed that several FTDP17-associated tau missense mutations made tau a more favorable substrate for abnormal hyperphosphorylation compared with wildtype tau. Both the phosphorylation kinetics, due to induced conformational changes, and the phosphorylation stoichiometry, due to increased phosphorylation of more than a single site, were more favorable in the mutant proteins. The mutant proteins polymerized into filaments more readily than wildtype tau, leading to decreased ability to bind wildtype tau.

Progressive Supranuclear Palsy

Conrad et al. (1997) demonstrated an association between progressive supranuclear palsy (PSNP; 601104) and a dinucleotide TG repeat polymorphism in intron 9 of the MAPT gene. They demonstrated overrepresentation of the most common allele (a0) and genotype (a0/a0) in PSNP. Baker et al. (1999) identified a series of polymorphisms scattered throughout the MAPT gene and described 2 extended haplotypes, designated H1 and H2, that cover the entire gene. The dinucleotide TG polymorphism alleles a0 (11 repeats), a1 (12 repeats), and a2 (13 repeats) are inherited with the H1 haplotype, whereas the a3 (14 repeats) and a4 (15 repeats) alleles are inherited with the H2 haplotype. In a total of approximately 200 unrelated Caucasian individuals, there was complete disequilibrium between polymorphisms that spanned the gene, suggesting that the establishment of the 2 haplotypes was an ancient event and that either recombination was suppressed in this region, or recombinant genes were selected against. Baker et al. (1999) showed that the more common haplotype, designated H1, was significantly overrepresented in patients with progressive supranuclear palsy, extending earlier reports of the association between the intronic dinucleotide polymorphism allele a0 and the disorder.

Litvan et al. (2001) examined 63 patients with PSNP and found that the presence of the tau H1/H1 genotype was significantly greater in patients compared to controls. There was no difference between PSNP cases with one H1 or two H1 alleles in the age of onset, severity, or survival of patients, thus showing that tau genotyping does not predict the prognosis of PSNP. However, Litvan et al. (2001) noted that most of the PSNP patients carried the H1/H1 genotype (88.9%) and none of the patients carried the H2/H2 genotype, thus limiting the conclusions of the study.

Using single-nucleotide polymorphisms, Pittman et al. (2004) mapped linkage disequilibrium (LD) in the regions flanking MAPT and established the maximum extent of the haplotype block on chromosome 17q21.31 as a region covering approximately 2 Mb. The gene-rich region extended centromerically beyond the corticotropin-releasing hormone receptor-1 gene (CRHR1; 122561) to a region of approximately 400 kb, where there was a complete loss of LD. The telomeric end was defined by an approximately 150-kb region just beyond the WNT3 (165330) gene. The authors showed that the entire, fully extended H1 haplotype was associated with PSNP, which implicates several other genes in addition to MAPT as candidate pathogenic loci.

Rademakers et al. (2005) and Pittman et al. (2005) used a large collection of pathologically confirmed PSNP samples to fine map PSNP risk on H1 chromosomes in PSNP cases and controls. PSNP risk was associated with an extended subhaplotype (H1c), and the risk for PSNP was narrowed to a 22-kb region in intron 0 of MAPT by examining younger patients with, presumably, a larger genetic component to their disease. The most likely explanation of the association of the MAPT H1 haplotype and PSNP is that variants in the H1 (and H2) haplotypes confer risk of (protect against) disease by altering expression at the locus, with the risky H1 haplotype expressing higher levels of MAPT.

Kwok et al. (2008) showed that the MAPT G-to-A allele of rs242557, which partially defines the H1c subhaplotype, results in increased MAPT gene expression.

Pick Disease

Zhukareva et al. (2002) used biochemical, immunohistochemical, and ultrastructural methods to characterize pathologic tau isoform composition in 14 sporadic Pick disease (172700) brains. They found that both 3R and 4R microtubule-binding isoforms were present in gray and white matter of various brain regions, particularly the cortex and hippocampus. Specifically, 7 cases had predominantly pathologic 3R isoforms, 4 cases had a mixture of 3R and 4R isoforms, and 3 cases had primarily 4R isoforms. Isolated tau filaments were primarily straight, but twisted forms were also present. Although the cases shared similar clinical and neuropathologic features, the biochemical profiles of abnormal tau were diverse.

Parkinson Disease

In a large study of 1,056 individuals from 235 families selected from 13 clinical centers in the United States and Australia and from a family ascertainment core center, Martin et al. (2001) found that haplotypes of single-nucleotide polymorphisms (SNPs) yielded strong evidence of association with late-onset Parkinson disease (PD; 168600). Positive association was found with 1 haplotype (P = 0.009) and a negative association with another haplotype (P = 0.007). The results were thought to implicate MAPT as a susceptibility gene for idiopathic Parkinson disease.

Kwok et al. (2004) identified several SNPs in the MAPT promoter region corresponding to the H1 and H2 haplotypes, as well as a novel variant of the H1 haplotype, termed H1-prime. In a cohort of 206 patients with idiopathic late-onset Parkinson disease, the authors found a significant association with the H1/H1 promoter genotype. In vitro analysis showed that the H1 haplotype was more efficient at driving gene expression than the H2 haplotype, suggesting that increased MAPT expression is a susceptibility factor in idiopathic PD.

Alternate mRNA splicing of exons 2, 3, and 10 of the MAPT gene results in the expression of 6 polypeptides in the human CNS (Higuchi et al., 2002). The predominant isoforms differ by the presence of either 3 or 4 microtubule-binding domains, the 3-repeat (3R) and 4-repeat (4R) isoforms, which result from the exclusion or inclusion of exon 10 (Panda et al., 2003). Frontotemporal dementia with parkinsonism linked to chromosome 17 (FTDP17; 600274) is called, in pathology terms, a '4R tauopathy' because of the presence of fibrillar aggregates comprising the 4R tau isoform. There are 2 predominant MAPT haplotypes--termed H1 and H2 and extending more than 500 kb--in which variants appear to be in complete linkage disequilibrium; H1 and H2 haplotypes do not recombine (Pastor et al., 2002). H1/H1 homozygous genotypes are overrepresented in 4R tauopathies. Using H1-specific SNPs, Skipper et al. (2004) demonstrated that MAPT H1 is a misnomer and consists of a family of recombining H1 alleles. Population genetics, linkage disequilibrium, and association analyses showed that specific MAPT H1 subhaplotypes are preferentially associated with Parkinson disease. Using a sliding scale of MAPT H1-specific haplotypes in age- and sex-matched PD cases and controls from central Norway, Skipper et al. (2004) refined the disease association to within an interval of approximately 90 kb of the 5-prime end of the MAPT locus.

In a study of 557 PD patient-control pairs, Mamah et al. (2005) found that individuals with the SNCA Rep1 261/261 or MAPT H1/H1 genotypes had an increased risk of PD compared to those with neither genotype (odds ratio of 1.96); however, the combined effect of the 2 genotypes was the same as for either genotype alone. Mamah et al. (2005) suggested that the MAPT H1/H1 genotype may cause increased SNCA fibrillization in persons with lower SNCA protein concentrations due to genotypes other than Rep1 261/261. In persons with the Rep1 261/261 genotype, the MAPT H1/H1 genotype confers no additional risk because the SNCA protein is already at threshold concentration for self-fibrillization.

Kwok et al. (2005) identified 2 functional SNPs in the GSK3B (605004) gene that influenced its transcriptional activity and correlated with enhanced phosphorylation of MAPT in vitro, respectively. Conditional logistic regression analysis of the genotypes of 302 Caucasian PD patients and 184 Chinese PD patients found an association between the GSK3B polymorphisms, MAPT haplotype, and risk of PD. Kwok et al. (2005) concluded that GSK3B polymorphisms interact with MAPT haplotypes to modify disease risk in PD. Garcia-Gorostiaga et al. (2009) confirmed the findings of Kwok et al. (2005) in a cohort of 314 Spanish patients with PD.

Among 1,762 PD patients, Zabetian et al. (2007) found a significant association between the H1/H1 genotype and risk of disease (odds ratio of 1.46; p = 8 x 10(-7)). The effect was evident in both familial and sporadic subgroups, men and women, and early- and late-onset disease. Within H1/H1 individuals, there was no association with H1 subhaplotypes.

Among 659 PD patients, Goris et al. (2007) found a synergistic interaction between the MAPT H1 haplotype and an A-to-G SNP (rs356219) in the 3-prime region of the SNCA gene. Carrying the combination of risk genotypes at both loci approximately doubled the risk of disease (p = 3 x 10(-6)). The findings suggested that MAPT and SNCA are involved in shared or converging pathogenic pathways and may have a synergistic effect. Cognitive decline and the development of dementia was associated with the H1/H1 genotype (p = 10(-4)). In a final analysis that combined data from other studies, Goris et al. (2007) confirmed the association of the H1/H1 genotype with PD (odds ratio of 1.4; p = 2 x 10(-19)).

In a study of 543 PD patients from 296 families with a proband and at least 1 affected first-degree relative as part of the GenePD Study group, Tobin et al. (2008) found a significant association between the H1 haplotype and PD (odds ratio of 1.72; p = 0.0008). In particular, rs1800547 of the H1 haplotype was significantly associated with PD (p = 0.02 after Bonferroni correction). Tobin et al. (2008) also identified a novel H1 subhaplotype that predicted an even greater increased risk for PD (OR, 4.48; p = 0.003), but some of these SNPs extended beyond the 3-prime region of MAPT. The expression of 4R MAPT, STH (607067), and KIAA1267 was significantly increased in cerebellum samples from PD patients relative to controls. No difference in expression was observed for 3R repeat MAPT. The findings suggested that variations in MAPT gene expression may underlie the association with PD.

Among 202 Spanish patients with PD, Seto-Salvia et al. (2011) found a significant association between the H1 haplotype and the development of dementia (odds ratio of 3.73, p = 0.002). Examination of subhaplotypes showed that a rare version of H1, named H1p, was overrepresented in PD patients with dementia compared to controls (2.3% vs 0.1%, p = 0.003). There was a protective effect for the H2a haplotype on PD with dementia. There was no association between MAPT variants and disease among 164 patients with Alzheimer disease or 41 with Lewy body dementia.

In a statistical analysis of 5,302 PD patients and 4,161 controls from 15 sites, Elbaz et al. (2011) found no evidence for an interactive effect between the H1 haplotype in the MAPT gene and SNPs in the SNCA gene on disease. Variation in each gene was associated with PD risk, indicating independent effects.

Alzheimer Disease

Conrad et al. (2002) identified a single-nucleotide polymorphism that results in a gly7-to-arg (G7R) change in the saitohin gene which appeared to be overrepresented in the homozygous state in late-onset Alzheimer disease subjects.

Because the MAPT R406W mutation (157140.0003) can cause a clinical picture closely resembling Alzheimer disease (see 104300) and quite different from frontotemporal dementia with parkinsonism, Rademakers et al. (2003) stated that MAPT should be considered a candidate gene for clinical Alzheimer disease families in which mutations in known Alzheimer disease genes have been excluded.

Myers et al. (2005) reported that the H1c subhaplotype of MAPT on the background of the well-described H1 clade was associated with risk of Alzheimer disease in 360 autopsy-confirmed cases with ages at death over 65 years of age and 252 controls.

Kwok et al. (2008) showed that the MAPT G-to-A allele of rs242557, which partially defines the H1c subhaplotype, results in increased MAPT gene expression. The authors also provided evidence that the H1/H2 MAPT haplotype interacts with functional SNPs in the GSK3B gene (605004) to affect risk of Alzheimer disease.


Genotype/Phenotype Correlations

Among 22 patients with FTLD (600274) due to a MAPT mutation, Whitwell et al. (2009) found different patterns of gray matter atrophy using MRI voxel-based morphometry. All patients showed gray matter loss in the anterior temporal lobes, with varying degrees of involvement of the frontal and parietal lobes. Within the temporal lobe, individuals with the IVS10+16 (157140.0006), IVS10+3, N279K (157140.0009), or S305N (157140.0010) mutations showed gray matter loss particularly affecting the medial temporal lobes, including the hippocampus and amygdala. These mutations are all predicted to influence the alternative splicing of MAPT pre-mRNA, resulting in increased 4R tau isoforms. In contrast, patients with the P301L (157140.0001) or V337M (157140.0008) mutations showed gray matter loss particularly affecting the inferior and lateral temporal lobes, with a relative sparing of the medial temporal lobe. P301L and V337M mutation carriers also showed gray matter loss in the basal ganglia. These mutations are predicted to affect the structure and functional properties of the tau protein, which are more prone to aggregation. The different patterns suggested a potential difference in mutant protein function resulting from different pathogenic mutations.


Animal Model

Stambolic et al. (1996) observed that lithium treatment of intact eukaryotic cells inhibits GSK3-dependent phosphorylation of the microtubule-associated protein tau, a putative GSK3 substrate.

Using Western blot analysis, Hiesberger et al. (1999) demonstrated that mice lacking either Reelin (RELN; 600514) or Vldlr (192977) and ApoER2 (LRP8; 602600) exhibit a dramatic increase in the phosphorylation level of the microtubule-stabilizing protein tau.

To model tauopathies, Ishihara et al. (1999) overexpressed the smallest human tau isoform in the central nervous system of transgenic mice. These mice acquired age-dependent central nervous system pathology similar to FTDP17, including insoluble, hyperphosphorylated tau and argyrophilic intraneuronal inclusions formed by tau-immunoreactive filaments. Inclusions were present in cortical and brainstem neurons but were most abundant in spinal cord neurons, where they were associated with axon degeneration, diminished microtubules, and reduced axonal transport in ventral roots, as well as spinal cord gliosis and motor weakness. These transgenic mice recapitulated key features of tauopathies and provided models for elucidating mechanisms underlying diverse tauopathies, including Alzheimer disease.

To model aspects of tau-related physiology and pathology, Spittaels et al. (1999) generated transgenic mice that overexpressed the 4R human tau isoform specifically in neurons. The mice developed axonal degeneration in brain and spinal cord, characterized by reduced or blocked axonal transport and axonal dilations as well as sensorimotor deficits, in the absence of intraneuronal neurofibrillary tangles. Spittaels et al. (1999) noted that excess normal tau might be sufficient to cause neuronal injury and suggested that excess of the 4R tau protein interferes with kinesin-dependent transport by saturating binding sites on microtubules. Spittaels et al. (2000) showed that when GSK3-beta (GSK3B; 605004) is expressed in these transgenic mice there is a strong reduction in the number of axonal dilations and a nearly complete alleviation of motor impairments, presumably by phosphorylation of tau, which then reduces the binding of tau to microtubules. Although more phosphorylated tau was available, neither PHFs nor tangles were formed.

Tesseur et al. (2000) observed that overexpression of human APOE4 in neurons of transgenic mice resulted in hyperphosphorylation of tau, and Tesseur et al. (2000) demonstrated that these mice exhibited widespread astrogliosis in the brain, impaired axonal transport, and axonal degeneration.

Lewis et al. (2000) demonstrated that expression of human tau containing the most common mutation, P301L (157140.0001), results in motor and behavioral deficits in transgenic mice, with age- and gene-dose-dependent development of neurofibrillary tangles (NFT). This phenotype occurred as early as 6.5 months in hemizygous and 4.5 months in homozygous animals. Abnormalities were seen not only in the central nervous system; a peripheral neuropathy and skeletal muscle involvement with neurogenic atrophy were also found.

Gotz et al. (2001) demonstrated that injection of beta-amyloid (104760) A-beta-42 fibrils into the brains of P301L mutant tau transgenic mice causes 5-fold increases in the numbers of neurofibrillary tangles in cell bodies within the amygdala from where neurons project to the injection sites. Gallyas silver impregnation identified neurofibrillary tangles that contained tau phosphorylated at serine-212/threonine-214 and serine-422. Neurofibrillary tangles were composed of twisted filaments and occurred in 6-month-old mice as early as 18 days after A-beta-42 injections. Gotz et al. (2001) concluded that their data support the hypothesis that A-beta-42 fibrils can accelerate neurofibrillary tangle formation in vivo.

Lewis et al. (2001) crossed JNPL3 transgenic mice expressing a mutant tau protein, which developed neurofibrillary tangles and progressive motor disturbance, with Tg2576 transgenic mice expressing mutant beta-amyloid precursor protein (APP), thus modulating the APP-A-beta environment. The resulting double-mutant (tau/APP) progeny and the Tg2576 parental strain developed amyloid-beta deposits at the same age; however, relative to JNPL3 mice, the double mutants exhibited neurofibrillary tangle pathology that was substantially enhanced in the limbic system and olfactory cortex. Lewis et al. (2001) concluded that either APP or amyloid-beta influences the formation of neurofibrillary tangles. The interaction between amyloid-beta and tau pathologies in these mice supports the hypothesis that a similar interaction occurs in Alzheimer disease.

Nguyen et al. (2001) found that hyperphosphorylation of neurofilament and tau proteins was associated with abnormal elevation of the p25/p35 (see 603460) ratio and Cdk5 (123831) activity in the spinal cord of a transgenic mouse model of amyotrophic lateral sclerosis (ALS; 105400).

Using proteomic analysis, David et al. (2005) showed that expression of human P301L mutant tau in transgenic mice resulted in distinct modifications of the brain proteome, suggesting alterations in the mitochondrial electron transport chain, cellular antioxidant capacities, and synaptic properties. Subsequent examination of complex V levels in brains of FTDP17 patients carrying the P301L tau mutation confirmed the observations made in P301L tau transgenic mice and suggested that P301L mutant tau pathology caused a specific mitochondrial dysfunction in humans and mice. In agreement, transgenic P301L tau mice exhibited an initial defect in mitochondrial function with reduced complex I activity, which, with age, translated into a mitochondrial respiration deficiency with diminished ATP synthesis corresponding to reduced complex V activity. P301L mutant tau also caused higher oxidative stress, modified lipid peroxidation levels, and upregulated antioxidant enzyme activities, without reducing mitochondrial numbers or significantly changing transport of mitochondria along neurites. In addition, P301L mutant tau decreased the membrane potential of cortical brain cells in transgenic mice, as these cells became more susceptible to A-beta treatment.

Taniguchi et al. (2005) found that transgenic mice expressing human N279K (157140.0009) mutant tau were viable, with normal feeding and body weight. However, transgenic mice displayed cognitive/sensorimotor deficits when evaluated by a comprehensive battery of tests.

SantaCruz et al. (2005) found that mice expressing a repressible human tau variant developed progressive age-related neurofibrillary tangles, neuronal loss, and behavioral impairments. After the suppression of transgenic tau, memory function recovered and neuron numbers stabilized, but neurofibrillary tangles continued to accumulate. SantaCruz et al. (2005) concluded that neurofibrillary tangles are not sufficient to cause cognitive decline or neuronal death in this model of tauopathy.

Karsten et al. (2006) identified the Npepps gene (606793) as a tau modifier by using a cross-species functional genomic approach to analyze gene expression in mice. Npepps expression was increased in multiple brain regions in a mouse model of frontotemporal dementia (FTD) compared to control mice. In Drosophila, Npepps protected against tau-induced neurodegeneration, whereas loss of Npepps exacerbated neurodegeneration. Immunoblot, SDS-PAGE, and Western blot analyses showed that human NPEPPS directly proteolyzed and significantly diminished human tau. Western blot analysis of 6 brains derived from human FTD patients showed increased NPEPPS expression, particularly in the cerebellum.

Yoshiyama et al. (2007) developed transgenic mice expressing wildtype human MAPT or MAPT with the pro301-to-ser (P301S; 157140.0012) mutation. P301S mice developed synaptic pathology and microgliosis in hippocampus at 3 months of age, followed by synaptic dysfunction at 6 months of age, prior to neuron loss and formation of neurofibrillary tangles. Yoshiyama et al. (2007) concluded that synaptic pathology and microgliosis may be the earliest manifestation of tauopathies.

Chatterjee et al. (2009) employed a Drosophila model of tauopathy to investigate the interdependence of tau kinases in regulating the phosphorylation and toxicity of tau in vivo. Tau mutants resistant to phosphorylation by Par1, the fly homolog of MARK1 (606511), were less toxic than wildtype tau; however, this was not due to their resistance to phosphorylation by Shaggy (GSK3B; 605004). On the contrary, a tau mutant resistant to phosphorylation by Shaggy retained substantial toxicity and had increased affinity for microtubules compared with wildtype tau. Chatterjee et al. (2009) suggested that, in addition to tau phosphorylation, microtubule binding may play a crucial role in the regulation of tau toxicity when misexpressed.

Expression of human tau in C. elegans neurons causes accumulation of aggregated tau leading to neurodegeneration and uncoordinated movement. Guthrie et al. (2009) used this model of human tauopathy disorders to screen for genes required for tau neurotoxicity. Recessive loss-of-function mutations in the Sut2 locus (ZC3H14; 613279) suppress the worm Unc phenotype, tau aggregation, and neurodegenerative changes caused by human tau. Guthrie et al. (2009) cloned the Sut2 gene and found that it encodes a novel subtype of CCCH zinc finger protein conserved across animal phyla. Sut2 shares significant identity with the mammalian SUT2. A yeast 2-hybrid screen revealed that Sut2 bound to Zyg12, the sole C. elegans HOOK protein family member (see 607820). Likewise, Sut2 bound Zyg12 in in vitro protein binding assays. Loss of Zyg12 led to a marked upregulation of Sut2 protein supporting the connection between Sut2 and Zyg12. The human ortholog of Sut2 bound only to HOOK2 (607824), suggesting that the interaction between Sut2 and HOOK family proteins may be conserved across animal phyla.

Iijima-Ando et al. (2010) showed that the DNA damage-activated checkpoint kinase-2 (CHK2; 604373) is a novel tau kinase. Overexpression of Drosophila Chk2 increased tau phosphorylation at ser262 and enhanced tau-induced neurodegeneration in transgenic flies expressing human tau. The nonphosphorylatable ser262-to-ala mutation abolished Chk2-induced enhancement of tau toxicity, suggesting that the ser262 phosphorylation site may be involved in the enhancement of tau toxicity by Chk2. In vitro kinase assays revealed that human CHK2 and a closely related checkpoint kinase, CHK1 (603078), directly phosphorylated human tau at ser262. Drosophila Chk2 did not modulate the activity of the fly homolog of microtubule affinity regulating kinase (see MARK3, 602678), which has been shown to be a physiologic tau ser262 kinase. Iijima-Ando et al. (2010) suggested that CHK1 and CHK2 may be involved in tau phosphorylation and toxicity in the pathogenesis of Alzheimer disease.

Using transgenic Drosophila expressing human A-beta-42 (APP; 104760) and tau, Iijima et al. (2010) showed that tau phosphorylation at ser262 plays a critical role in A-beta-42-induced tau toxicity. Coexpression of A-beta-42 increased tau phosphorylation at AD-related sites including ser262, and enhanced tau-induced neurodegeneration. In contrast, formation of either sarkosyl-insoluble tau or paired helical filaments was not induced by A-beta-42. Coexpression of A-beta-42 and tau carrying the nonphosphorylatable ser262ala mutation did not cause neurodegeneration, suggesting that the ser262 phosphorylation site is required for the pathogenic interaction between A-beta-42 and tau. CHK2 phosphorylates tau at ser262 and enhances tau toxicity in a transgenic Drosophila model (Iijima-Ando et al., 2010). Exacerbation of A-beta-42-induced neuronal dysfunction by blocking tumor suppressor p53 (191170), a key transcription factor for the induction of DNA repair genes, in neurons suggested that induction of a DNA repair response is protective against A-beta-42 toxicity. The authors concluded that tau phosphorylation at ser262 is crucial for A-beta-42-induced tau toxicity in vivo, and they suggested a model of AD progression in which activation of DNA repair pathways is protective against A-beta-42 toxicity but may trigger tau phosphorylation and toxicity in AD pathogenesis.

Using Drosophila and mouse models of tauopathies, Falzone et al. (2010) showed that reductions in axonal transport by reduction in kinesin-1 (KLC1; 600025) expression can exacerbate human tau protein hyperphosphorylation, formation of insoluble aggregates, and tau-dependent neurodegeneration. The authors hypothesized that nonlethal reductions in axonal transport, and perhaps other types of minor axonal stress, are sufficient to induce and/or accelerate abnormal tau behavior characteristic of Alzheimer disease and other neurodegenerative tauopathies.

Xu et al. (2010) generated transgenic mice overexpressing full-length human TTBK1 and tau with the P301L mutation. Transgenic mice developed age-dependent pathology in central nervous system, including intraneuronal accumulation of phosphorylated tau, accumulation of sarkosyl-soluble phospho-tau multimers, Cdk5 and Gsk3-beta activation, locomotor dysfunction, and motor neuron degeneration. The results suggested that TTBK1 is involved in phosphorylation-dependent generation of pathogenic tau aggregation.

Lei et al. (2012) found decreased levels of soluble tau and increased iron levels in the substantia nigra of postmortem samples from patients with Parkinson disease compared to controls. The tau loss was independent of neuronal loss. Similar changes were observed in the MPTP mouse model of Parkinson disease. Mapt-knockout mice developed age-dependent brain atrophy, iron accumulation in various brain regions, and neuronal loss in the substantia nigra, as well as cognitive deficits and parkinsonism. These changes could be prevented by oral treatment with a moderate iron chelator, clioquinol. In primary neuronal cell cultures, Lei et al. (2012) found that loss of tau caused iron retention by decreasing the surface expression of APP, which interacts with ferroportin (SLC40A1; 604653) to accomplish neuronal iron efflux. The findings suggested that tau is needed to prevent age-related damage, that loss of soluble tau contributes to toxic neuronal iron accumulation in Alzheimer disease, Parkinson disease, and other tauopathies, and that pharmacologic intervention in this process may ameliorate the disease.

Taylor et al. (2018) showed that transgenic C. elegans expressing the kinase catalytic domain of human TTBK1 or TTBK2 (611695) were behaviorally normal. However, C. elegans coexpressing TTBK1 or TTBK2 with tau caused neurodegeneration, behavioral abnormalities, aberrant phosphorylation, and shortened lifespan. Coexpression of TTBK2 with high levels of tau resulted in lethality.

Przybyla et al. (2020) found that transgenic mice expressing human tau with a pro301-to-ser (P301S) mutation displayed progressive spatial learning deficits accompanied by reduced synaptic plasticity and aberrant neuronal network activity in brain. Mutant mice presented with an immediate early gene response, consistent with increased neuronal activity. Increased immediate early gene activity was confined to neurons harboring tau pathology, providing a cellular link between aberrant tau and network dysfunction.


History

Neve et al. (1986) noted that 1 of their tau clones hybridized to both chromosome 17 and to an additional region of homology on chromosome 6q21 (MAPTL). They suggested that either the 2 clones represented 2 different tau genes on different chromosomes, or that the 2 clones encoded different regions of the same expressed gene on chromosome 17 and only one of them bore homology to a nonexpressed pseudogene on chromosome 6.

Volz et al. (1994) reported that MAPTL clustered with FTHP1 and GSTA1 (138359) in the region 6p21.1-p12, with the map order unknown.


ALLELIC VARIANTS 25 Selected Examples):

.0001   DEMENTIA, FRONTOTEMPORAL, WITH PARKINSONISM

SUPRANUCLEAR PALSY, PROGRESSIVE, 1, INCLUDED
MAPT, PRO301LEU
SNP: rs63751273, gnomAD: rs63751273, ClinVar: RCV000015313, RCV000084527, RCV000763405, RCV002508757, RCV003407335

In a large Dutch kindred with frontotemporal dementia (600274) reported by Heutink et al. (1997), Hutton et al. (1998) found a pro301-to-leu mutation (P301L) in exon 10 of the MAPT gene. The same mutation was found in a small kindred from the United States. This substitution occurred in a highly conserved region of the MAPT sequence, where a proline residue was found in all mammalian species from which tau had been cloned to that time. The P301L mutation would affect only the 4-repeat tau isoforms because exon 10 is spliced out of mRNA that encodes the 3-repeat isoforms. Analysis of tau aggregates in affected brains from the U.S. kindred revealed that these consist mainly of 4-repeat isoforms, consistent with the mutation affecting exon 10.

In a note added in proof, Poorkaj et al. (1998) described finding a C-to-T transition at nucleotide 728 of the MAPT gene in a newly ascertained family with FTDP17. The mutation resulted in a PRO243LEU mutation. This is the same mutation as that reported by Hutton et al. (1998), who used a different numbering system for the nucleotides and codons. At the time the paper of Poorkaj et al. (1998) was submitted, the longest amino acid form of tau in the database did not include all of the alternatively spliced exons (Poorkaj, 1998).

MAPT transcripts that contain this exon 10 mutation encode tau isoforms with 4 microtubule (MT)-binding repeats (4Rtau) as opposed to tau isoforms with 3 MT-binding repeats (3Rtau). Clark et al. (1998) found that brains of patients with the P301L missense mutation contained aggregates of insoluble 4Rtau in filamentous inclusions, which may lead to neurodegeneration.

Using purified recombinant proteins, Alonso et al. (2004) showed that several FTDP17-associated tau mutations, including P301L, made tau a more favorable substrate for abnormal hyperphosphorylation compared with wildtype tau. Both the phosphorylation kinetics, due to induced conformational changes, and the phosphorylation stoichiometry, due to increased phosphorylation of more than a single site, were more favorable in the mutant proteins. The mutant proteins polymerized into filaments more readily than wildtype tau, leading to decreased ability to bind wildtype tau.

Using proteomic analysis, David et al. (2005) showed that expression of human P301L mutant tau in transgenic mice resulted in distinct modifications of the brain proteome, suggesting alterations in the mitochondrial electron transport chain, cellular antioxidant capacities, and synaptic properties. Subsequent examination of complex V levels in brains of FTDP17 patients carrying the P301L tau mutation confirmed the observations made in P301L tau transgenic mice and suggested that P301L mutant tau pathology caused a specific mitochondrial dysfunction in humans and mice. In agreement, transgenic P301L tau mice exhibited an initial defect in mitochondrial function with reduced complex I activity, which, with age, translated into a mitochondrial respiration deficiency with diminished ATP synthesis corresponding to reduced complex V activity. P301L mutant tau also caused higher oxidative stress, modified lipid peroxidation levels, and upregulated antioxidant enzyme activities, without reducing mitochondrial numbers or significantly changing transport of mitochondria along neurites. In addition, P301L mutant tau decreased the membrane potential of cortical brain cells in transgenic mice, as these cells became more susceptible to A-beta treatment.

Using purified recombinant proteins, Aoyagi et al. (2007) showed that P301L mutant tau assembled into nuclei more rapidly than wildtype tau or R406W (157140.0003) mutant tau. However, P301L mutant nuclei could only promote assembly of P301L mutant tau into filaments, whereas wildtype and R406W mutant nuclei had the ability to seed both wildtype and P301L mutant tau. Pronase digestion experiments revealed conformational differences between P301L mutant tau and wildtype or R406W mutant tau. The core structure of P301L mutant tau seeds was distinct from that of wildtype tau seeds, regardless of phosphorylation state, whereas R406W mutant tau seeds had a core structure similar to that of wildtype tau seeds.

Donker Kaat et al. (2009) identified a P301L mutation in 1 of 172 probands with progressive supranuclear palsy (601104).


.0002   DEMENTIA, FRONTOTEMPORAL, WITH PARKINSONISM

MAPT, GLY272VAL
SNP: rs63750376, ClinVar: RCV000015315, RCV000084519

In a second large Dutch kindred with frontotemporal dementia (600274) reported by Heutink et al. (1997) as an example of hereditary Pick disease, Hutton et al. (1998) found a gly272-to-val mutation (G272V) that affected a highly conserved residue within the microtubule-associated domain, encoded by exon 9 of the MAPT gene.

Using purified recombinant proteins, Alonso et al. (2004) showed that several FTDP17-associated tau mutations, including G272V, made tau a more favorable substrate for abnormal hyperphosphorylation compared with wildtype tau. Both the phosphorylation kinetics, due to induced conformational changes, and the phosphorylation stoichiometry, due to increased phosphorylation of more than a single site, were more favorable in the mutant proteins. The mutant proteins polymerized into filaments more readily than wildtype tau, leading to decreased ability to bind wildtype tau.


.0003   DEMENTIA, FRONTOTEMPORAL

MAPT, ARG406TRP
SNP: rs63750424, gnomAD: rs63750424, ClinVar: RCV000015316, RCV000084554, RCV002476970

In affected members of a family from the United States with autosomal dominant frontotemporal dementia (600274) originally reported by Reed et al. (1997), Hutton et al. (1998) identified a heterozygous arg406-to-trp mutation (R406W) in the MAPT gene. Neuropathologic examination identified tau-positive intraneuronal neurofibrillary tangles similar to those found in Alzheimer disease (see 104300).

Connell et al. (2001) studied the in vitro phosphorylation of several tau mutants by glycogen synthase kinase 3-beta and confirmed previous findings that the R406W mutation had the surprising effect of reducing tau phosphorylation in cells, compared to both the wildtype and other mutant forms. Connell et al. (2001) suggested that reduced tau phosphorylation and a minor loss of function associated with R406W may contribute to the later onset and somewhat milder phenotype found in families with this mutation.

Tatebayashi et al. (2002) showed that the expression of the R406W mutation in transgenic mice resulted in the development of congophilic hyperphosphorylated tau inclusions in forebrain neurons. These inclusions appeared as early as 18 months of age. As with human cases, tau inclusions were composed of both mutant and endogenous wildtype tau, and were associated with microtubule disruption and flame-shaped transformations of the affected neurons. Behaviorally, aged transgenic mice had associative memory impairment without obvious sensorimotor deficits. Therefore, these mice exhibited a phenotype mimicking R406W FTDP17 (frontotemporal dementia and parkinsonism linked to chromosome 17).

Saito et al. (2002) reported a patient who presented at age 47 years with psychiatric disturbances, primarily delusions, who developed overt dementia by age 52, and died at age 53. There was rapid progression in the last year of life. His father had had a similar illness. Postmortem examination revealed neuronal loss associated with neurofibrillary tangles and neuropil threads, accentuated in the medial temporal lobe, that were immunoreactive for the tau protein. Molecular analysis revealed an R406W mutation.

Rademakers et al. (2003) suggested that the R406W mutation can cause a clinical picture closely resembling Alzheimer disease and quite different from frontotemporal dementia with parkinsonism. They described a 6-generation Belgian family that carried the R406W mutation and had such clinical features, and pointed to reports of 2 other families that carried the R406W mutation and had similar clinical characteristics: 1 from the U.S. Midwest with a Danish ancestor (Reed et al., 1997) and 1 from the Netherlands (van Swieten et al., 1999). Haplotype data ruled against a founder effect for the origin of the mutation in western Europe. Rademakers et al. (2003) stated that their report illustrated the phenotypic heterogeneity of MAPT mutations and reemphasized that MAPT should be considered a candidate gene for clinical Alzheimer disease families in which mutations in known Alzheimer disease genes have been excluded.

Using purified recombinant proteins, Alonso et al. (2004) showed that several FTDP17-associated tau mutations, including R406W, made tau a more favorable substrate for abnormal hyperphosphorylation compared with wildtype tau. Both the phosphorylation kinetics, due to induced conformational changes, and the phosphorylation stoichiometry, due to increased phosphorylation of more than a single site, were more favorable in the mutant proteins. The mutant proteins polymerized into filaments more readily than wildtype tau, leading to decreased ability to bind wildtype tau.

Using purified recombinant proteins, Aoyagi et al. (2007) showed that P301L (157140.0001) mutant tau assembled into nuclei more rapidly than wildtype tau or R406W mutant tau. However, P301L mutant nuclei could only promote assembly of P301L mutant tau into filaments, whereas wildtype and R406W mutant nuclei had the ability to seed both wildtype and P301L mutant tau. Pronase digestion experiments revealed conformational differences between P301L mutant tau and wildtype or R406W mutant tau. The core structure of P301L mutant tau seeds was distinct from that of wildtype tau seeds, regardless of phosphorylation state, whereas R406W mutant tau seeds had a core structure similar to that of wildtype tau seeds.


.0004   DEMENTIA, FRONTOTEMPORAL, WITH PARKINSONISM

MAPT, IVS10, C-U, +14
SNP: rs63750972, gnomAD: rs63750972, ClinVar: RCV000015317, RCV000084536

In addition to 3 missense mutations, Hutton et al. (1998) found 3 heterozygous mutations in a cluster of 4 nucleotides 13- to 16-bp 3-prime of the exon 10 5-prime splice site of the MAPT gene in families with frontotemporal dementia with parkinsonism (600274). Six families had mutations at 1 of these 3 sites, including 4 families in which the disorder had previously been linked to chromosome 17. One of these linked families was reported by Wilhelmsen et al. (1994) as disinhibition-dementia-parkinsonism-amyotrophy complex (DDPAC; 600274). In this family, the mRNA showed a C-to-U transition in the stem-loop structure at position 14 in the splice donor site of intron 10.


.0005   DEMENTIA, FRONTOTEMPORAL, WITH PARKINSONISM

MAPT, IVS10, A-G, +13
SNP: rs63750308, ClinVar: RCV000015318, RCV000084535

One of the splice site mutations identified by Hutton et al. (1998) in families with frontotemporal dementia with parkinsonism (600274) was an A-to-G transition at position 13 in intron 10.

In a British patient with frontotemporal dementia, Pickering-Brown et al. (2002) identified the tau IVS10 +13 mutation. He presented with apathy and inertia, and later developed semantic loss.


.0006   DEMENTIA, FRONTOTEMPORAL, WITH PARKINSONISM

MAPT, IVS10, C-T, +16
SNP: rs63751011, ClinVar: RCV000084537, RCV000626752, RCV000626753, RCV000687510, RCV003415858

In an Australian family with frontotemporal dementia with parkinsonism (600274), Hutton et al. (1998) found a C-to-T change in the MAPT mRNA at position 16 of the splice donor site of intron 10. The mutation resulted in an increased incorporation of exon 10 in MAPT mRNA levels, which increased the proportion of tau isoforms containing 4 microtubule-binding domains.

Goedert et al. (1999) identified a heterozygous IVS10+16C-T mutation in the MAPT gene in affected members of a family with frontotemporal dementia and extrapyramidal signs. Initial neuropathologic examination (Lanska et al., 1994) showed prominent subcortical gliosis (221820), but later studies by Goedert et al. (1999) showed hyperphosphorylated tau in both neurons and glial cells with wide twisted ribbons made of 4-repeat tau isoforms. The molecular findings confirmed the diagnosis. Goedert et al. (1999) noted that phenotypic heterogeneity associated with MAPT mutations has led to classification of related diseases into distinct entities.

Janssen et al. (2002) described the clinical characteristics of 9 families with frontotemporal dementia and the tau exon 10 +16 mutation and found considerable variation in age at onset and duration of disease both between and within families, suggesting the influence of other genetic or environmental factors.

Pickering-Brown et al. (2002) reported 8 British families with frontotemporal dementia from North Wales in which affected members had the tau exon 10 +16 mutation. Clinical features included disinhibition, restless overactivity, fatuous affect, puerile behavior, verbal and motor stereotypes, and semantic loss. In 5 of these 8 families, Pickering-Brown et al. (2004) found a common 3-cM haplotype flanked by markers D17S1860 and D17S806. An affected Australian pedigree (Dark, 1997), 7 patients from London (Janssen et al., 2002; Lantos et al., 2002), and 2 patients from Philadelphia (Poorkaj et al., 2001) had the same haplotype. The disease haplotype was not observed in 100 geographically matched control chromosomes. Pickering-Brown et al. (2004) concluded that the exon 10 +16 mutation represents a founder effect that originated in North Wales. They demonstrated that the mutation is on the H1 tau haplotype.

Doran et al. (2007) reported a large family from Liverpool, England, in which 8 individuals had frontotemporal dementia associated with the IVS10+16C-T mutation. All patients were initially diagnosed with Alzheimer disease (104300) because of presentation of memory deficits and word-finding difficulties. Prototypic features of frontotemporal dementia, such as disinhibition and personality changes, were not noted initially. Doran et al. (2007) noted the phenotypic variability of this mutation.

Colombo et al. (2009) noted that the identification of the IVS10+16C-T mutation in patients from North America and the U.K. support the hypothesis of a founder effect of British origin. Using several different methods of haplotype analysis, the authors estimated that the mutation occurred about 23 generations ago, around 1300 A.D., before Welsh people started emigrating to the U.S. and Australia, where they introduced the mutation.


.0007   DEMENTIA, FRONTOTEMPORAL, WITH PARKINSONISM

MAPT, IVS10, G-A, +1
SNP: rs1568327531, ClinVar: RCV000015320

In a kindred known to have presenile dementia identified as multiple system tauopathy with presenile dementia (MSTD; 600274) in individuals in 5 successive generations, Spillantini et al. (1998) found a G-to-A transition in the nucleotide 3-prime of the exon 10 splice donor site. It was identified in 11 affected family members and segregated with the disease haplotype in other family members. Examination of the nucleotide sequence of exon 10 and the 5-prime intron junction identified a predicted stem-loop structure that encompassed the last 6 nucleotides at the 3-prime end of exon 10 and 19 nucleotides of the intron, including the GT splice donor site. The G-to-A transition destabilized this stem-loop structure. The disorder was associated with an abnormal preponderance of soluble tau protein isoforms with 4 microtubule-binding repeats over isoforms with 3 repeats. This was thought most likely to account for the previous finding that sarkosyl-insoluble tau protein extracted from the filamentous deposits in familial MSTD consists only of tau isoforms with 4 repeats. The departure from the normal ratio of 4-repeat to 3-repeat tau isoforms leads to the formation of abnormal tau filaments. The dysregulation of tau protein production in turn leads to neurodegeneration.


.0008   DEMENTIA, FRONTOTEMPORAL, WITH PARKINSONISM

MAPT, VAL337MET
SNP: rs63750570, ClinVar: RCV000015321, RCV000084548

Poorkaj et al. (1998) identified 9 DNA sequence variants in 2 families with what they referred to as FTDP17 (600274); 8 of these were also identified in controls and were thus considered polymorphisms. The ninth variant, VAL279MET, was found in 1 FTDP17 family, but not in the other. (This mutation is designated val337 to met in the numbering used by Hutton et al. (1998); see 157140.0001.) The change was considered positive since it occurred in a highly conserved residue and a normal valine is found at this position in all 3 tau interrepeat sequences and in other MAPT homologs. Furthermore, the mutation cosegregated with the disease in the family and was not found in normal controls. At the time that the paper of Poorkaj et al. (1998) was submitted, the longest amino acid form of tau in the database did not include all of the alternatively spliced exons (Poorkaj, 1998).

Using purified recombinant proteins, Alonso et al. (2004) showed that several FTDP17-associated tau mutations, including V337M, made tau a more favorable substrate for abnormal hyperphosphorylation compared with wildtype tau. Both the phosphorylation kinetics, due to induced conformational changes, and the phosphorylation stoichiometry, due to increased phosphorylation of more than a single site, were more favorable in the mutant proteins. The mutant proteins polymerized into filaments more readily than wildtype tau, leading to decreased ability to bind wildtype tau.

Sohn et al. (2019) found that the FTD-associated tau V337M mutant shortened the axon initial segment (AIS) and impaired AIS plasticity in human induced pluripotent stem cell (iPSC)-derived neurons. Electrophysiologic properties of tau V337M neurons revealed that the mutation also impaired homeostatic control of spontaneous neuronal activity in response to depolarization. End-binding protein-3 (EB3, or MAPRE3; 605788), a component of the AIS cytoskeleton, was associated with ANKG (106280) and tau in the AIS submembrane region of human neurons. However, the V337M mutation increased the binding affinity of tau with EB3, leading to increased EB3 accumulation in the AIS in tau V337M mutant neurons. EB3 accumulation increased its interaction with ANKG and promoted its structural stability via physical interaction with microtubules, thereby impairing AIS plasticity in tau V337M mutant neurons.


.0009   DEMENTIA, FRONTOTEMPORAL, WITH PARKINSONISM

MAPT, ASN279LYS
SNP: rs63750756, gnomAD: rs63750756, ClinVar: RCV000015322, RCV000084521, RCV000763404

In affected members of a kindred with pallidopontonigral degeneration (PPND; 600274) originally reported by Wszolek et al. (1992), Clark et al. (1998) identified an asn279-to-lys (N279K) mutation in exon 10 of the MAPT gene. Although the clinical features and associated regional variations in the neuronal loss observed in different kindreds with frontotemporal dementia and parkinsonism linked to chromosome 17 are diverse, the diagnostic lesions in the brain are tau-rich filaments in the cytoplasm of specific subpopulations of neurons and glial cells. The insoluble tau aggregates isolated from brains of patients with the N279K mutation were analyzed by immunoblotting using tau-specific antibodies. For each of 3 mutations, abnormal tau with an apparent relative mass of 64 and 69 kD was observed. The dephosphorylated material comigrated with tau isoforms containing exon 10 having 4 microtubule-binding repeats but not with 3-repeat tau. Thus, the brains contained aggregates of insoluble 4Rtau in filamentous inclusions, which may lead to neurodegeneration.

Delisle et al. (1999) reported 2 French brothers with the N279K mutation who presented early in the fourth decade with a neurodegenerative disorder characterized by an akinetic rigid syndrome and dementia. There was widespread neuronal and glial tau accumulation in the cortex, basal ganglia, brainstem nuclei, and white matter.

Arima et al. (2000) reported 2 Japanese brothers with the N279K mutation who presented with frontotemporal dementia characterized by personality changes, behavioral disinhibition, asocial misconduct, recent memory deficits, and parkinsonism. The disease progressed in both patients, rendering them mute and bedridden with death at ages 57 and 50. Pathologic examination showed severe temporal lobe atrophy, neuronal loss, and tau-immunoreactive neurofibrillary tangles and cytoplasmic inclusions. There was also phosphorylated tau deposition in neurons of the spinal cord and degeneration of the lateral corticospinal tracts. Insoluble tau, mainly 4-repeat isoforms, with molecular masses of 64 and 69 kD were observed. The tau aggregates were found to be composed of paired hollow tubules, 11 to 12 nm in diameter. One of the patients had responded temporarily to L-DOPA therapy.

Yasuda et al. (1999) reported a Japanese patient with the N279K mutation who presented at 41 years of age with rapidly progressive parkinsonism which later included dementia, hallucinations, vertical gaze palsy, extensor plantar responses, frontal release signs, and incontinence. His sister had parkinsonism and dementia and died in her forties. Neuropathologic examination of the proband showed severe cortical atrophy, discoloration of the striatum, globus pallidus, and luysian body, and prominent depigmentation of the substantia nigra and locus ceruleus. In both cortical and subcortical regions, there was moderate to severe neuronal loss, astrocytic gliosis, tau-positive neuronal inclusion bodies, and loss of motor neurons. The authors termed this disease pallido-nigro-luysian degeneration. Wszolek et al. (2000) suggested that the disorder in the patient reported by Yasuda et al. (1999), in which parkinsonism was the initial prominent feature, is a subtype of FTDP17.

Tsuboi et al. (2002) analyzed clinical and genealogic records of 4 previously reported families with the N279K mutation: an American family (Clark et al., 1998), a French family (Delisle et al., 1999), and 2 Japanese families (Yasuda et al., 1999; Arima et al., 2000). The clinical phenotype was similar in all families: onset in the forties, survival time of 6 to 8 years, parkinsonism as the presenting sign, short or no response to L-DOPA treatment, progressive dementia, pyramidal dysfunction, and gaze palsy. Molecular genetic studies showed a shared disease haplotype between the 2 Japanese families, suggesting a founder effect. Tsuboi et al. (2002) suggested that FTDP17 could be divided into 2 major clinical groups, parkinsonism-predominant and dementia-predominant, and that patients with the N279K mutation fall into the former category. Tsuboi et al. (2002) also reported a previously undescribed Japanese patient with the N279K mutation and a similar phenotype.

Rademakers et al. (2004) likewise commented on the highly similar clinical phenotypes associated with the N279K mutation despite the occurrence on different genetic backgrounds. In the various families, the mean age at onset ranged from 41 to 47 years. In patients, typical parkinsonian features such as bradykinesia, rigidity, and postural instability were uniformly seen. Personality and behavioral changes and dementia also occurred during the course of the illness, but were less prominent.

Taniguchi et al. (2005) found that transgenic mice expressing human N279K mutant tau were viable, with normal feeding and body weight. However, transgenic mice displayed cognitive/sensorimotor deficits when evaluated by a comprehensive battery of tests.


.0010   DEMENTIA, FRONTOTEMPORAL, WITH PARKINSONISM

MAPT, SER305ASN
SNP: rs63751165, ClinVar: RCV000015323, RCV000084530

Iijima et al. (1999) reported a Japanese family with early-onset hereditary frontotemporal dementia (600274) and a novel ser305-to-asn mutation in the tau gene. The patients presented with personality changes followed by impaired cognition and memory as well as disorientation, but minimal parkinsonism.


.0011   PICK DISEASE

MAPT, GLY389ARG
SNP: rs63750512, gnomAD: rs63750512, ClinVar: RCV000015324, RCV000517183, RCV001851871

Murrell et al. (1999) described a gly389-to-arg (G389R) missense mutation in exon 13 of the MAPT gene in a patient with a condition closely resembling Pick disease (172700). When 38 years old, the proband presented with progressive aphasia and memory disturbance, followed by apathy, indifference, and hyperphagia. MRI showed the dramatic progression of cerebral atrophy. PET revealed marked glucose hypometabolism that was most severe in the left frontal, temporal, and parietal cortical regions. Rigidity, pyramidal signs, and profound dementia progressed until death at 43 years of age. A paternal uncle, who had died at 43 years of age, had presented with similar symptoms. The proband's brain showed numerous tau-immunoreactive Pick body-like inclusions.

Of 30 cases of pathologically confirmed Pick disease, Pickering-Brown et al. (2000) identified 2 mutations in the tau gene in 2 unrelated patients: a G-to-A change, resulting in a G389R substitution, and an A-to-C change, resulting in a lys257-to-thr substitution (157140.0014). The patient with the G389R mutation showed a decline in intellectual ability with forgetfulness, aggression, and a decline in personal hygiene at age 32, which progressed to death by age 37. Pathologic examination showed severe atrophy and neuronal loss in the frontal cortex with many tau-immunoreactive neuronal inclusions. In vitro, the G389R mutation reduced the ability of tau to promote microtubule assembly by 25 to 30%.


.0012   DEMENTIA, FRONTOTEMPORAL, WITH PARKINSONISM

MAPT, PRO301SER
SNP: rs63751438, ClinVar: RCV000015325, RCV000084526

In affected members of a family with early-onset frontotemporal dementia with parkinsonism (600274) and seizures, Sperfeld et al. (1999) identified a 1137C-T transition in exon 10 of the MAPT gene, resulting in a pro301-to-ser (P301S) substitution.

In a father and son with frontotemporal dementia and corticobasal degeneration, respectively, Bugiani et al. (1999) identified a P301S substitution in the MAPT gene. Both individuals developed rapidly progressive disease in the third decade. Neuropathic examination of the father showed extensive filamentous pathology made of hyperphosphorylated tau protein. Recombinant P301S tau protein showed a greatly reduced ability to promote microtubule assembly.

Yasuda et al. (2000) identified a P301S mutation in the MAPT gene in a Japanese man with onset of parkinsonism at age 37 and rapidly progressive frontotemporal dementia beginning at age 39. Citing earlier reports of mutations at the same position, the authors noted that codon 301 of the tau gene is a hotspot of pathogenic mutations and that the mutations exhibit variable phenotypes (see P301L; 157140.0001).

Lossos et al. (2003) identified the P301S mutation in 3 affected members of a Jewish family of Algerian origin with rapidly progressive frontotemporal dementia with parkinsonism. Disease onset was characterized by personality changes in the late thirties, followed by progressive cognitive and motor deterioration leading to akinetic mutism or death within 3 to 5 years. Werber et al. (2003) identified the P301S mutation in a 39-year-old Jewish woman of Algerian origin with frontotemporal dementia with parkinsonism. Family history revealed 4 affected first-degree relatives, 1 of whom was still alive and also carried the mutation. The families reported by Lossos et al. (2003) and Werber et al. (2003) were members of an extended Israeli-French kindred (Yasuda et al., 2005).

Yasuda et al. (2005) reported another Japanese family in which 6 members had frontotemporal dementia and parkinsonism caused by a heterozygous P301S mutation. Age at disease onset ranged from 28 to 40 years, and clinical findings were characterized by parkinsonism in all patients with dementia reported only in the 3 younger patients who had detailed medical histories. Neuropsychiatric features of these 3 patients included emotional incontinence, euphoria, apathy, and perseveration. One patient had psychotic features. Neuropathologic examination of these 3 patients showed neuronal loss and gliosis most prominent in the substantia nigra, globus pallidus, and subthalamic nucleus associated with neuropil thread-rich, tau-containing lesions. Yasuda et al. (2005) noted the phenotypic variability associated with the P301S mutation.


.0013   DEMENTIA, FRONTOTEMPORAL

MAPT, ASN296ASN
SNP: rs63750912, ClinVar: RCV000015326, RCV000084525

In a family with frontotemporal dementia (600274), previously reported by Brown et al. (1996) as having corticobasal degeneration, Spillantini et al. (2000) identified a silent mutation in exon 10 of the tau gene, resulting in exon trapping, or the inclusion of multiple copies of exon 10 in the mRNA transcript. This mutation was thought to lead to an abnormal preponderance of soluble tau protein isoforms with 4 microtubule-binding repeats over isoforms with 3 repeats, perhaps by disrupting a cis-acting element in the region of exon 10. Affected members demonstrated progressive dementia beginning in the fifth or sixth decade with extensive tau pathology and frontotemporal atrophy.


.0014   DEMENTIA, FRONTOTEMPORAL, WITH PARKINSONISM

MAPT, GLU342VAL
SNP: rs63750711, ClinVar: RCV000015327, RCV000084549

In a patient with familial frontotemporal dementia (600274), Lippa et al. (2000) identified a 1025A-T transversion in exon 12 of the tau gene, resulting in a glu342-to-val (E342V) substitution. The authors found that the levels of the 4R0N isoform were increased in all brain regions, whereas the 4RN1 and 4RN2 isoforms were decreased, a previously undescribed pathologic tau profile. Lippa et al. (2000) suggested that exon 12 may influence alternative splicing of other exons within the tau gene. Pathologic examination revealed frontotemporal neuron loss, intracytoplasmic tau aggregates, and paired helical tau filaments.


.0015   PICK DISEASE

MAPT, LYS257THR
SNP: rs63750129, gnomAD: rs63750129, ClinVar: RCV000015328, RCV000084515

Of 30 cases of pathologically confirmed Pick disease (172700), Pickering-Brown et al. (2000) identified 2 mutations in the tau gene in 2 unrelated patients: an A-to-C change, resulting in a lys257-to-thr substitution, and the previously identified G389R substitution(157140.0011). Tau-immunoreactive Pick bodies and Pick cells were present. In vitro, the K257T mutation reduced the ability of tau to promote microtubule assembly by 70%.


.0016   PICK DISEASE

MAPT, LYS369ILE
SNP: rs63751264, ClinVar: RCV000015329, RCV000084552

In a patient with Pick disease (172700) characterized clinically by onset at age 52 of rapidly progressing decline of cognitive and behavioral abilities, and pathologically by severe temporal atrophy and the presence of Pick inclusion bodies and Pick cells, Neumann et al. (2001) identified a mutation in the tau gene, resulting in a lys369-to-ile (K339I) substitution in exon 12. The K369I mutation led to a 90% reduction in the rate of microtubule assembly, and the authors suggested that free mutant tau may assemble abnormally, leading to pathologic changes.


.0017   DEMENTIA, FRONTOTEMPORAL, WITH PARKINSONISM

MAPT, ARG5HIS
SNP: rs63750959, gnomAD: rs63750959, ClinVar: RCV000015330, RCV000266864

In a patient with late-onset (age 75 years) frontotemporal dementia with parkinsonism (600274), Hayashi et al. (2002) identified a G-to-A transition in exon 1 of the MAPT gene, resulting in an arg5-to-his (R5H) substitution. Pathologic examination revealed frontotemporal atrophy, neuronal loss, widespread tau-immunoreactive glial cytoplasmic inclusions, and insoluble tau filaments composed of 4-repeat tau. The mutation reduced the ability of tau to promote microtubule assembly and promoted fibril formation in vitro. The patient had an elderly brother with dementia who died at age 86 years.


.0018   PICK DISEASE

MAPT, SER320PHE
SNP: rs63750635, ClinVar: RCV000015331, RCV000084544, RCV000995804

In a patient who presented with presenile dementia characterized by mild memory problems at age 38 years, which progressed to major memory problems, personality changes, and aphasia at age 47, followed by death at age 53, Rosso et al. (2002) identified a C-to-T transition in exon 11 of the MAPT gene, resulting in a ser320-to-phe (S320F) substitution. Postmortem examination revealed findings consistent with Pick disease (172700), including focal bilateral atrophy of the anterior temporal lobes, extensive tau pathology in the form of Pick-like bodies, and insoluble tau-immunoreactive filaments. In vitro studies showed that the mutation resulted in a markedly reduced ability of tau to promote microtubule assembly.


.0019   SUPRANUCLEAR PALSY, PROGRESSIVE, 1

MAPT, ARG5LEU
SNP: rs63750959, gnomAD: rs63750959, ClinVar: RCV000084498, RCV002508758

Among 96 patients with progressive supranuclear palsy (601104), Poorkaj et al. (2002) identified 1 patient with a G-to-T mutation in a highly conserved position in exon 1 of the tau gene, resulting in an arg5-to-leu (R5L) substitution. Functional studies showed that the mutation delayed assembly initiation and lowered the mass of microtubules formed, but the assembly rate was increased compared to normal tau. The authors hypothesized a gain-of-function mutation. (Replacement of arginine by histidine at the same position (157170.0017) causes frontotemporal dementia with parkinsonism.)


.0020   DEMENTIA, FRONTOTEMPORAL, WITH PARKINSONISM

MAPT, IVS10, T-C, +11
SNP: rs63751394, ClinVar: RCV000015333, RCV000084533

In a Japanese patient with onset of FTDP17 (600274) at age 48 years and mental retardation since his first decade, Miyamoto et al. (2001) identified a heterozygous IVS10+11T-C change in the MAPT gene. An older brother, 1 sister, his mother, and grandfather had similar features, including mental retardation, but all had died and thus could not be tested. It was not clear if the mental retardation was related.


.0021   SUPRANUCLEAR PALSY, PROGRESSIVE, 1, ATYPICAL

PARKINSON DISEASE, LATE-ONSET, SUSCEPTIBILITY TO, INCLUDED
MAPT, ASN296DEL
SNP: rs63751392, ClinVar: RCV000015334, RCV000015335, RCV000084581

In a Spanish patient with atypical progressive supranuclear palsy (260540), born from a third-degree consanguineous marriage, Pastor et al. (2001) identified a homozygous 3-bp deletion (AAT) of asparagine at codon 296 in the MAPT gene. The proband and his brother demonstrated a remarkably similar phenotype characterized by onset in the late thirties of mild cognitive decline, inappropriate behavior, ocular movement abnormalities, and asymmetric parkinsonism. The parents were both heterozygous for the mutation, consistent with autosomal recessive inheritance. Two uncles, who were heterozygous for the mutation, developed late-onset typical Parkinson disease (168600). However, there were several asymptomatic heterozygous individuals over the age of 60 in the family, which the authors attributed to reduced penetrance. Pastor et al. (2001) noted that codon 296 lies in the region of the gene which alters splicing of exon 10 (see N296N; 157140.0013).

In a family with a variable neurodegenerative phenotype, including a progressive supranuclear palsy-like syndrome and parkinsonism, Rossi et al. (2004) identified a heterozygous deletion of the last 2 bases of codon 296 and the first base of codon 297, resulting in the deletion of asn296. The proband developed antecollis, dysarthria, postural instability with falls, slowing of ocular movements, and increased deep tendon reflexes at age 36 years, consistent with atypical PSNP. The mutation was also identified in a paternal aunt with typical dopa-responsive Parkinson disease, in 2 asymptomatic sisters of the proband, and in 3 asymptomatic daughters of a deceased paternal uncle who had atypical dopa-unresponsive Parkinson disease with pyramidal signs and cognitive impairment. The authors suggested incomplete penetrance of the disorder.

Oliva and Pastor (2004) noted that although the nucleotide changes reported by Rossi et al. (2004) were unique, the resultant amino acid change, deletion of asn296, is the same as that reported by Pastor et al. (2001); therefore the mutation should be considered the same. Following international recommendations (den Dunnen and Antonarakis, 2000), Oliva and Pastor (2004) proposed that the complete nomenclature for this mutation be designated 'c713-715delATA' at the cDNA level and 'delN296' at the protein level. Oliva and Pastor (2004) noted that functional studies had shown that the delN296 change can decrease the ability of 4-repeat tau to promote microtubule assembly (Yoshida et al., 2002). The authors suggested that heterozygosity for the delN296 mutation may be a risk factor for both a PSNP-like syndrome and Parkinson disease.


.0022   DEMENTIA, FRONTOTEMPORAL

MAPT, LEU266VAL
SNP: rs63750349, gnomAD: rs63750349, ClinVar: RCV000015336, RCV000084517

In 2 brothers with frontotemporal dementia (600274) characterized by onset of personality changes in the thirties with progression to dementia, Kobayashi et al. (2003) identified a heterozygous C-to-G transversion in exon 9 of the MAPT gene, resulting in a leu266-to-val (L266V) substitution. Their mother reportedly had a similar disease, but was not examined. The mutation was not present in 200 control subjects. In addition, both brothers were homozygotes for the H1 tau haplotype, which has been shown to be overrepresented in patients with FTD compared to controls. Brain neuropathologic examination in one brother showed frontotemporal atrophy with severe neuronal loss and gliosis in both the cortices and the substantia nigra. There were also tau-positive neuronal inclusions and tau-positive astrocytes with stout filaments. Recombinant tau proteins with the L266V mutation were less able to promote microtubule assembly than wildtype tau.


.0023   SUPRANUCLEAR PALSY, PROGRESSIVE, 1

MAPT, SER352LEU
SNP: rs63750425, ClinVar: RCV000084550, RCV002508759

In 2 sibs from a consanguineous marriage who presented with a form of progressive supranuclear palsy (601104) characterized by fatal respiratory hypoventilation, Nicholl et al. (2003) identified a homozygous 1291C-T transition in exon 12 of the MAPT gene, resulting in a nonconserved ser352-to-leu (S352L) substitution in the N-terminal repeat of the tau protein. The authors called the disorder 'tauopathy and respiratory failure.' The 29-year-old pregnant sister developed dyspnea with stridor, and later had a generalized seizure that left her unconscious. Despite therapy, she died after 9 days. Her 30-year-old brother developed cough syncope, dyspnea, and central apnea, which progressed over 40 months, leading to death at age 34 years. During the illness, he showed slow smooth pursuit and impaired saccades, mild rigidity and bradykinesia, and myoclonic jerks. Neuropathologic examination showed widespread neuronal eosinophilia and pyknosis with gliosis in multiple brain regions, consistent with hypoxic brain damage. There was pervasive tau pathology in neuronal perikarya, neurites, and threads in the gray matter of the hippocampus, thalamus, and pons, but not in the cerebral cortex. Functional studies indicated that the mutated protein showed reduced binding to microtubules as well as increased fibrillization and aggregation. Both sibs carried the H1/H1 haplotype associated with PSP, Nicholl et al. (2003) commented on the unusual apparent autosomal recessive inheritance of this tauopathy. In a review of the role of tau in neurodegenerative diseases, Quadros et al. (2007) classified the disorder in the sibs reported by Nicholl et al. (2003) as PSP.


.0024   DEMENTIA, FRONTOTEMPORAL, WITH PARKINSONISM

MAPT, LYS317MET
SNP: rs63750092, ClinVar: RCV000015338, RCV000084543

In affected members of 2 families from the Basque country in Spain with a tauopathy best described as frontotemporal dementia with parkinsonism (600274), Zarranz et al. (2005) identified a heterozygous A-to-T transversion in exon 11 of the MAPT gene, resulting in a lys317-to-met (K317M) substitution. Mean age at disease onset was 48 years, characterized by dysarthria and features of parkinsonism. All patients developed parkinsonism and a pyramidal syndrome, and half had amyotrophy. Other variable features included supranuclear gaze palsy, bulbar palsy, and dystonia. Although behavioral changes were not a prominent feature, most patients had frontal signs, and cognitive decline appeared late in the disease. All patients eventually became fully dependent, wheelchair- or bed-bound, tetraplegic, mute, and unable to feed orally. Neuropathologic examination showed frontal and temporal lobe atrophy, severe neuronal loss in the substantia nigra, and diffuse patchy neuronal loss and gliosis with numerous tau- and ubiquitin-positive neurons. Six of 7 spinal cords examined showed severe neuronal loss in the anterior horn and degeneration of the corticospinal tracts, similar to that seen in amyotrophic lateral sclerosis. Two bands of phospho-tau of 68 and 64 kD were observed in brain tissue from 1 patient. Zarranz et al. (2005) emphasized the motor component of the syndrome in these patients. Haplotype analysis suggested a common origin in the 2 families, and genealogic analysis identified a probable common ancestor born in 1782.


.0025   SUPRANUCLEAR PALSY, PROGRESSIVE, 1

MAPT, GLY303VAL
SNP: rs63751391, ClinVar: RCV000084529, RCV002508760

In a patient with progressive supranuclear palsy (601104), Ros et al. (2005) identified a heterozygous 2095G-T transversion in exon 10 of the MAPT gene, resulting in a gly303-to-val (G303V) substitution. Three asymptomatic family members, who were younger than the average age at disease onset, also carried the mutation. Two deceased affected family members were obligate carriers of the mutation. Protein analysis of tissue from the proband's brain detected hyperphosphorylated tau protein and overexpression of tau isoforms with 4 microtubule-binding repeats. The mutation occurred in a highly conserved residue of the protein and was not identified in 194 control chromosomes.


REFERENCES

  1. Abel, K. J., Boehnke, M., Prahalad, M., Ho, P., Flejter, W. L., Watkins, M., VanderStoep, J., Chandrasekharappa, S. C., Collins, F. S., Glover, T. W., Weber, B. L. A radiation hybrid map of the BRCA1 region of chromosome 17q12-q21. Genomics 17: 632-641, 1993. [PubMed: 8244380] [Full Text: https://doi.org/10.1006/geno.1993.1383]

  2. Alonso, A. D. C., Grundke-Iqbal, I., Iqbal, K. Alzheimer's disease hyperphosphorylated tau sequesters normal tau into tangles of filaments and disassembles microtubules. Nature Med. 2: 783-787, 1996. [PubMed: 8673924] [Full Text: https://doi.org/10.1038/nm0796-783]

  3. Alonso, A. D., Mederlyova, A., Novak, M., Grundke-Iqbal, I., Iqbal, K. Promotion of hyperphosphorylation by frontotemporal dementia tau mutations. J. Biol. Chem. 279: 34873-34881, 2004. [PubMed: 15190058] [Full Text: https://doi.org/10.1074/jbc.M405131200]

  4. Andreadis, A., Brown, W. M., Kosik, K. S. Structure and novel exons of the human tau gene. Biochemistry 31: 10626-10633, 1992. [PubMed: 1420178] [Full Text: https://doi.org/10.1021/bi00158a027]

  5. Aoyagi, H., Hasegawa, M., Tamaoka, A. Fibrillogenic nuclei composed of P301L mutant tau induce elongation of P301L tau but not wild-type tau. J. Biol. Chem. 282: 20309-20318, 2007. [PubMed: 17526496] [Full Text: https://doi.org/10.1074/jbc.M611876200]

  6. Arima, K., Kowalska, A., Hasegawa, M., Mukoyama, M., Watanabe, R., Kawai, M., Takahashi, K., Iwatsubo, T., Tabira, T., Sunohara, N. Two brothers with frontotemporal dementia and parkinsonism with an N279K mutation of the tau gene. Neurology 54: 1787-1795, 2000. [PubMed: 10802785] [Full Text: https://doi.org/10.1212/wnl.54.9.1787]

  7. Baker, M., Litvan, I., Houlden, H., Adamson, J., Dickson, D., Perez-Tur, J., Hardy, J., Lynch, T., Bigio, E., Hutton, M. Association of an extended haplotype in the tau gene with progressive supranuclear palsy. Hum. Molec. Genet. 8: 711-715, 1999. [PubMed: 10072441] [Full Text: https://doi.org/10.1093/hmg/8.4.711]

  8. Brown, J., Lantos, P. L., Roques, P., Fidani, L., Rossor, M. N. Familial dementia with swollen achromatic neurons and corticobasal inclusion bodies: a clinical and pathological study. J. Neurol. Sci. 135: 21-30, 1996. [PubMed: 8926492] [Full Text: https://doi.org/10.1016/0022-510x(95)00236-u]

  9. Bugiani, O., Murrell, J. R., Giaccone, G., Hasegawa, M., Ghigo, G., Tabaton, M., Morbin, M., Primavera, A., Carella, F., Solaro, C., Grisoli, M., Savoiardo, M., Spillantini, M. G., Tagliavini, F., Goedert, M., Ghetti, B. Frontotemporal dementia and corticobasal degeneration in a family with a P301S mutation in tau. J. Neuropath. Exp. Neurol. 58: 667-677, 1999. [PubMed: 10374757] [Full Text: https://doi.org/10.1097/00005072-199906000-00011]

  10. Chatterjee, S., Sang, T.-K., Lawless, G. M., Jackson, G. R. Dissociation of tau toxicity and phosphorylation: role of GSK-3-beta, MARK and Cdk5 in a Drosophila model. Hum. Molec. Genet. 18: 164-177, 2009. [PubMed: 18930955] [Full Text: https://doi.org/10.1093/hmg/ddn326]

  11. Clark, L. N., Poorkaj, P., Wszolek, Z., Geschwind, D. H., Nasreddine, Z. S., Miller, B., Li, D., Payami, H., Awert, F., Markopoulou, K., Andreadis, A., D'Souza, I., Lee, V. M.-Y., Reed, L., Trojanowski, J. Q., Zhukareva, V., Bird, T., Schellenberg, G., Wilhelmsen, K. C. Pathogenic implications of mutations in the tau gene in pallido-ponto-nigral degeneration and related neurodegenerative disorders linked to chromosome 17. Proc. Nat. Acad. Sci. 95: 13103-13107, 1998. [PubMed: 9789048] [Full Text: https://doi.org/10.1073/pnas.95.22.13103]

  12. Colombo, R., Tavian, D., Baker, M. C., Richardson, A. M. T., Snowden, J. S., Neary, D., Mann, D. M. A., Pickering-Brown, S. M. Recent origin and spread of a common Welsh MAPT splice mutation causing frontotemporal lobar degeneration. Neurogenetics 10: 313-318, 2009. [PubMed: 19365643] [Full Text: https://doi.org/10.1007/s10048-009-0189-x]

  13. Connell, J. W., Gibb, G. M., Betts, J. C., Blackstock, W. P., Gallo, J.-M., Lovestone, S., Hutton, M., Anderton, B. H. Effects of FTDP-17 mutations on the in vitro phosphorylation of tau by glycogen synthase kinase 3-beta identified by mass spectrometry demonstrate certain mutations exert long-range conformational changes. FEBS Lett. 493: 40-44, 2001. [PubMed: 11278002] [Full Text: https://doi.org/10.1016/s0014-5793(01)02267-0]

  14. Conrad, C., Andreadis, A., Trojanowski, J. Q., Dickson,D. W., Kang, D., Chen, X., Weiderholt, W., Hansen, L., Masliah, E., Thal, L. J., Katzman, R., Xia, Y., Saitoh, T. Genetic evidence for the involvement of tau in progressive supranuclear palsy. Ann. Neurol. 41: 277-281, 1997. [PubMed: 9029080] [Full Text: https://doi.org/10.1002/ana.410410222]

  15. Conrad, C., Vianna, C., Freeman, M., Davies, P. A polymorphic gene nested within an intron of the tau gene: implications for Alzheimer's disease. Proc. Nat. Acad. Sci. 99: 7751-7756, 2002. [PubMed: 12032355] [Full Text: https://doi.org/10.1073/pnas.112194599]

  16. Dark, F. A family with autosomal dominant non-Alzheimer's presenile dementia. Aust. New Zeal. J. Psychiat. 31: 139-144, 1997. [PubMed: 9088499] [Full Text: https://doi.org/10.3109/00048679709073812]

  17. David, D. C., Hauptmann, S., Scherping, I., Schuessel, K., Keil, U., Rizzu, P., Ravid, R., Drose, S., Brandt, U., Muller, W. E., Eckert, A., Gotz, J. Proteomic and functional analyses reveal a mitochondrial dysfunction in P301L tau transgenic mice. J. Biol. Chem. 280: 23802-23814, 2005. [PubMed: 15831501] [Full Text: https://doi.org/10.1074/jbc.M500356200]

  18. de Calignon, A., Fox, L. M., Pitstick, R., Carlson, G. A., Bacskai, B. J., Spires-Jones, T. L., Hyman, B. T. Caspase activation precedes and leads to tangles. Nature 464: 1201-1204, 2010. [PubMed: 20357768] [Full Text: https://doi.org/10.1038/nature08890]

  19. Delacourte, A., Sergeant, N., Champain, D., Wattez, A., Maurage, C.-A., Lebert, F., Pasquier, F., David, J.-P. Nonoverlapping but synergetic tau and APP pathologies in sporadic Alzheimer's disease. Neurology 59: 398-407, 2002. [PubMed: 12177374] [Full Text: https://doi.org/10.1212/wnl.59.3.398]

  20. Delisle, M.-B., Murrell, J. R., Richardson, R., Trofatter, J. A., Rascol, O., Soulages, X., Mohr, M., Calvas, P., Ghetti, B. A mutation at codon 279 (N279K) in exon 10 of the tau gene causes a tauopathy with dementia and supranuclear palsy. Acta Neuropath. 98: 62-77, 1999. [PubMed: 10412802] [Full Text: https://doi.org/10.1007/s004010051052]

  21. den Dunnen, J. T., Antonarakis, S. E. Mutation nomenclature extensions and suggestions to describe complex mutations: a discussion. Hum. Mutat. 15: 7-12, 2000. Note: Erratum: Hum. Mutat. 20: 403 only, 2002. [PubMed: 10612815] [Full Text: https://doi.org/10.1002/(SICI)1098-1004(200001)15:1<7::AID-HUMU4>3.0.CO;2-N]

  22. Dixit, R., Ross, J. L., Goldman, Y. E., Holzbaur, E. L. F. Differential regulation of dynein and kinesin motor proteins by tau. Science 319: 1086-1089, 2008. [PubMed: 18202255] [Full Text: https://doi.org/10.1126/science.1152993]

  23. Donker Kaat, L., Boon, A. J. W., Azmani, A., Kamphorst, W., Breteler, M. M. B., Anar, B., Heutink, P., van Swieten, J. C. Familial aggregation of parkinsonism in progressive supranuclear palsy. Neurology 73: 98-105, 2009. [PubMed: 19458322] [Full Text: https://doi.org/10.1212/WNL.0b013e3181a92bcc]

  24. Donlon, T. A., Harris, P., Neve, R. L. Localization of microtubule-associated protein tau (MTBT1) to chromosome 17q21. (Abstract) Cytogenet. Cell Genet. 46: 607, 1987.

  25. Donnelly, M. P., Paschou, P., Grigorenko, E., Gurwitz, D., Mehdi, S. Q., Kajuna, S. L. B., Barta, C., Kungulilo, S., Karoma, N. J., Lu, R.-B., Zhukova, O. V., Kim, J.-J., and 11 others. The distribution and most recent common ancestor of the 17q21 inversion in humans. Am. J. Hum. Genet. 86: 161-171, 2010. [PubMed: 20116045] [Full Text: https://doi.org/10.1016/j.ajhg.2010.01.007]

  26. Doran, M., du Plessis, D. G., Ghadiali, E. J., Mann, D. M. A., Pickering-Brown, S., Larner, A. J. Familial early-onset dementia with tau intron 10 +16 mutation with clinical features similar to those of Alzheimer disease. Arch. Neurol. 64: 1535-1539, 2007. [PubMed: 17923640] [Full Text: https://doi.org/10.1001/archneur.64.10.1535]

  27. Elbaz, A., Ross, O. A., Ioannidis, J. P. A., Soto-Ortolaza, A. I., Moisan, F., Aasly, J., Annesi, G., Bozi, M., Brighina, L., Chartier-Harlin, M.-C., Destee, A., Ferrarese, C., and 29 others. Independent and joint effects of the MAPT and SNCA genes in Parkinson disease. Ann. Neurol. 69: 778-792, 2011. [PubMed: 21391235] [Full Text: https://doi.org/10.1002/ana.22321]

  28. Falcon, B., Zhang, W., Murzin, A. G., Murshudov, G., Garringer, H. J., Vidal, R., Crowther, R. A., Ghetti, B., Scheres, S. H. W., Goedert, M. Structures of filaments from Pick's disease reveal a novel tau protein fold. Nature 561: 137-140, 2018. [PubMed: 30158706] [Full Text: https://doi.org/10.1038/s41586-018-0454-y]

  29. Falcon, B., Zivanov, J., Zhang, W., Murzin, A. G., Garringer, H. J., Vidal, R., Crowther, R. A., Newell, K. L., Ghetti, B., Goedert, M., Scheres, S. H. W. Novel tau filament fold in chronic traumatic encephalopathy encloses hydrophobic molecules. Nature 568: 420-423, 2019. [PubMed: 30894745] [Full Text: https://doi.org/10.1038/s41586-019-1026-5]

  30. Falzone, T. L., Gunawardena, S., McCleary, D., Reis, G. F., Goldstein, L. S. B. Kinesin-1 transport reductions enhance human tau hyperphosphorylation, aggregation and neurodegeneration in animal models of tauopathies. Hum. Molec. Genet. 19: 4399-4408, 2010. [PubMed: 20817925] [Full Text: https://doi.org/10.1093/hmg/ddq363]

  31. Faraco, G., Hochrainer, K., Segarra, S. G., Schaeffer, S., Santisteban, M. M., Menon, A., Jiang, H., Holtzman, D. M., Anrather, J., Iadecola, C. Dietary salt promotes cognitive impairment through tau phosphorylation. Nature 574: 686-690, 2019. Note: Erratum: Nature 578: E9, 2019. [PubMed: 31645758] [Full Text: https://doi.org/10.1038/s41586-019-1688-z]

  32. Garcia-Gorostiaga, I., Sanchez-Juan, P., Mateo, I., Rodriguez-Rodriguez, E., Sanchez-Quintana, C., Curiel del Olmo, S., Vazquez-Higuera, J. L., Berciano, J., Combarros, O., Infante, J. Glycogen synthase kinase-3 and tau genes interact in Parkinson's and Alzheimer's diseases. (Letter) Ann. Neurol. 65: 759-760, 2009. [PubMed: 19557862] [Full Text: https://doi.org/10.1002/ana.21687]

  33. Giasson, B. I., Forman, M. S., Higuchi, M., Golbe, L. I., Graves, C. L., Kotzbauer, P. T., Trojanowski, J. Q., Lee, V. M.-Y. Initiation and synergistic fibrillization of tau and alpha-synuclein. Science 300: 636-640, 2003. [PubMed: 12714745] [Full Text: https://doi.org/10.1126/science.1082324]

  34. Goedert, M., Crowther, R. A., Spillantini, M. G. Tau mutations cause frontotemporal dementias. Neuron 21: 955-958, 1998. [PubMed: 9856453] [Full Text: https://doi.org/10.1016/s0896-6273(00)80615-7]

  35. Goedert, M., Spillantini, M. G., Crowther, R. A., Chen, S. G., Parchi, P., Tabaton, M., Lanska, D. J., Markesbery, W. R., Wilhelmsen, K. C., Dickson, D. W., Petersen, R. B., Gambetti, P. Tau gene mutation in familial progressive subcortical gliosis. Nature Med. 5: 454-457, 1999. [PubMed: 10202939] [Full Text: https://doi.org/10.1038/7454]

  36. Goedert, M., Spillantini, M. G., Jakes, R., Rutherford, D., Crowther, R. A. Multiple isoforms of human microtubule-associated protein tau: sequences and localization in neurofibrillary tangles of Alzheimer's disease. Neuron 3: 519-526, 1989. [PubMed: 2484340] [Full Text: https://doi.org/10.1016/0896-6273(89)90210-9]

  37. Goedert, M., Spillantini, M. G., Potier, M. C., Ulrich, J., Crowther, R. A. Cloning and sequencing of the cDNA encoding an isoform of microtubule-associated protein tau containing four tandem repeats: differential expression of tau protein mRNAs in human brain. EMBO J. 8: 393-399, 1989. [PubMed: 2498079] [Full Text: https://doi.org/10.1002/j.1460-2075.1989.tb03390.x]

  38. Goedert, M., Wischik, C. M., Crowther, R. A., Walker, J. E., Klug, A. Cloning and sequencing of the cDNA encoding a core protein of the paired helical filament of Alzheimer disease: identification as the microtubule-associated protein tau. Proc. Nat. Acad. Sci. 85: 4051-4055, 1988. [PubMed: 3131773] [Full Text: https://doi.org/10.1073/pnas.85.11.4051]

  39. Goode, B. L., Chau, M., Denis, P. E., Feinstein, S. C. Structural and functional differences between 3-repeat and 4-repeat tau isoforms: implications for normal tau function and the onset of neurodegenerative disease. J. Biol. Chem. 275: 38182-38189, 2000. [PubMed: 10984497] [Full Text: https://doi.org/10.1074/jbc.M007489200]

  40. Goris, A., Williams-Gray, C. H., Clark, G. R., Foltynie, T., Lewis, S. J. G., Brown, J., Ban, M., Spillantini, M. G., Compston, A., Burn, D. J., Chinnery, P. F., Barker, R. A., Sawcer, S. J. Tau and alpha-synuclein in susceptibility to, and dementia in, Parkinson's disease. Ann. Neurol. 62: 145-153, 2007. [PubMed: 17683088] [Full Text: https://doi.org/10.1002/ana.21192]

  41. Gotz, J., Chen, F., van Dorpe, J., Nitsch, R. M. Formation of neurofibrillary tangles in P301L tau transgenic mice induced by A-beta42 fibrils. Science 293: 1491-1495, 2001. [PubMed: 11520988] [Full Text: https://doi.org/10.1126/science.1062097]

  42. Guo, J.-P., Arai, T., Miklossy, J., McGeer, P. L. Amyloid-beta and tau form soluble complexes that may promote self aggregation of both into the insoluble forms observed in Alzheimer's disease. Proc. Nat. Acad. Sci. 103: 1953-1958, 2006. [PubMed: 16446437] [Full Text: https://doi.org/10.1073/pnas.0509386103]

  43. Guthrie, C. R., Schellenberg, G. D., Kraemer, B. C. SUT-2 potentiates tau-induced neurotoxicity in Caenorhabditis elegans. Hum. Molec. Genet. 18: 1825-1838, 2009. [PubMed: 19273536] [Full Text: https://doi.org/10.1093/hmg/ddp099]

  44. Hayashi, S., Toyoshima, Y., Hasegawa, M., Umeda, Y., Wakabayashi, K., Tokiguchi, S., Iwatsubo, T., Takahashi, H. Late-onset frontotemporal dementia with a novel exon 1 (Arg5His) tau gene mutation. Ann. Neurol. 51: 525-530, 2002. [PubMed: 11921059] [Full Text: https://doi.org/10.1002/ana.10163]

  45. Heutink, P., Stevens, M., Rizzu, P., Bakker, E., Kros, J. M., Tibben, A., Niermeijer, M. F., van Duijn, C. M., Oostra, B. A., van Swieten, J. C. Hereditary frontotemporal dementia is linked to chromosome 17q21-q22: a genetic and clinicopathological study of three Dutch families. Ann. Neurol. 41: 150-159, 1997. [PubMed: 9029063] [Full Text: https://doi.org/10.1002/ana.410410205]

  46. Heutink, P. Untangling tau-related dementia. Hum. Molec. Genet. 9: 979-986, 2000. [PubMed: 10767321] [Full Text: https://doi.org/10.1093/hmg/9.6.979]

  47. Hiesberger, T., Trommsdorff, M., Howell, B. W., Goffinet, A., Mumby, M. C., Cooper, J. A., Herz, J. Direct binding of reelin to VLDL receptor and apoE receptor 2 induces tyrosine phosphorylation of disabled-1 and modulates tau phosphorylation. Neuron 24: 481-489, 1999. [PubMed: 10571241] [Full Text: https://doi.org/10.1016/s0896-6273(00)80861-2]

  48. Higuchi, M., Lee, V. M.-Y., Trojanowski, J. Q. Tau and axonopathy in neurodegenerative disorders. Neuromolec. Med. 2: 131-150, 2002. [PubMed: 12428808] [Full Text: https://doi.org/10.1385/NMM:2:2:131]

  49. Holzer, M., Craxton, M., Jakes, R., Arendt, T., Goedert, M. Tau gene (MAPT) sequence variation among primates. Gene 341: 313-322, 2004. [PubMed: 15474313] [Full Text: https://doi.org/10.1016/j.gene.2004.07.013]

  50. Hong, M., Zhukareva, V., Vogelsberg-Ragaglia, V., Wszolek, Z., Reed, L., Miller, B. I., Geschwind, D. H., Bird, T. D., McKeel, D., Goate, A., Morris, J. C., Wilhelmsen, K. C., Schellenberg, G. D., Trojanowski, J. Q., Lee, V. M.-Y. Mutation-specific functional impairments in distinct tau isoforms of hereditary FTDP-17. Science 282: 1914-1917, 1998. [PubMed: 9836646] [Full Text: https://doi.org/10.1126/science.282.5395.1914]

  51. Hutton, M., Lendon, C. L., Rizzu, P., Baker, M., Froelich, S., Houlden, H., Pickering-Brown, S., Chakraverty, S., Isaacs, A., Grover, A., Hackett, J., Adamson, J., and 39 others. Association of missense and 5-prime-splice-site mutations in tau with the inherited dementia FTDP-17. Nature 393: 702-705, 1998. [PubMed: 9641683] [Full Text: https://doi.org/10.1038/31508]

  52. Hutton, M. Missense and splice site mutations in tau associated with FTDP-17: multiple pathogenic mechanisms. Neurology 56 (suppl. 4): S21-S25, 2001. [PubMed: 11402146] [Full Text: https://doi.org/10.1212/wnl.56.suppl_4.s21]

  53. Iijima, K., Gatt, A., Iijima-Ando, K. Tau ser262 phosphorylation is critical for A-beta-42-induced tau toxicity in a transgenic Drosophila model of Alzheimer's disease. Hum. Molec. Genet. 19: 2947-2957, 2010. [PubMed: 20466736] [Full Text: https://doi.org/10.1093/hmg/ddq200]

  54. Iijima, M., Tabira, T., Poorkaj, P., Schellenberg, G. D., Trojanowski, J. Q., Lee, V. M., Schmidt, M. L., Takahashi, K., Nabika, T., Matsumoto, T., Yamashita, Y., Yoshioka, S., Ishino, H. A distinct familial presenile dementia with a novel missense mutation in the tau gene. Neuroreport 10: 497-501, 1999. [PubMed: 10208578] [Full Text: https://doi.org/10.1097/00001756-199902250-00010]

  55. Iijima-Ando, K., Zhao, L., Gatt, A., Shenton, C., Iijima, K. A DNA damage-activated checkpoint kinase phosphorylates tau and enhances tau-induced neurodegeneration. Hum. Molec. Genet. 19: 1930-1938, 2010. [PubMed: 20159774] [Full Text: https://doi.org/10.1093/hmg/ddq068]

  56. Ishihara, T., Hong, M., Zhang, B., Nakagawa, Y., Lee, M. K., Trojanowski, J. Q., Lee, V. M.-Y. Age-dependent emergence and progression of a tauopathy in transgenic mice overexpressing the shortest human tau isoform. Neuron 24: 751-762, 1999. [PubMed: 10595524] [Full Text: https://doi.org/10.1016/s0896-6273(00)81127-7]

  57. Janssen, J. C., Warrington, E. K., Morris, H. R., Lantos, P., Brown, J., Revesz, T., Wood, N., Khan, M. N., Cipolotti, L., Fox, N. C., Rossor, M. N. Clinical features of frontotemporal dementia due to the intronic tau 10 +16 mutation. Neurology 58: 1161-1168, 2002. [PubMed: 11971081] [Full Text: https://doi.org/10.1212/wnl.58.8.1161]

  58. Jiang, Z., Cote, J., Kwon, J. M., Goate, A. M., Wu, J. Y. Aberrant splicing of tau pre-mRNA caused by intronic mutations associated with the inherited dementia frontotemporal dementia with Parkinsonism linked to chromosome 17. Molec. Cell. Biol. 20: 4036-4048, 2000. Note: Erratum: Molec. Cell. Biol. 20: 5360 only, 2000. [PubMed: 10805746] [Full Text: https://doi.org/10.1128/MCB.20.11.4036-4048.2000]

  59. Karsten, S. L., Sang, T.-K., Gehman, L. T., Chatterjee, S., Liu, J., Lawless, G. M., Sengupta, S., Berry, R. W., Pomakian, J., Oh, H. S., Schulz, C., Hui, K.-S., Wiedau-Pazos, M., Vinters, H. V., Binder, L. I., Geschwind, D. H., Jackson, G. R. A genomic screen for modifiers of tauopathy identified puromycin-sensitive aminopeptidase as an inhibitor of tau-induced neurodegeneration. Neuron 51: 549-560, 2006. [PubMed: 16950154] [Full Text: https://doi.org/10.1016/j.neuron.2006.07.019]

  60. Kobayashi, T., Ota, S., Tanaka, K., Ito, Y., Hasegawa, M., Umeda, Y., Motoi, Y., Takanashi, M., Yasuhara, M., Anno, M., Mizuno, Y., Mori, H. A novel L266V mutation of the tau gene causes frontotemporal dementia with a unique tau pathology. Ann. Neurol. 53: 133-137, 2003. [PubMed: 12509859] [Full Text: https://doi.org/10.1002/ana.10447]

  61. Kondo, A., Shahpasand, K., Mannix, R., Qiu, J., Moncaster, J., Chen, C.-H., Yao, Y., Lin, Y.-M., Driver, J. A., Sun, Y., Wei, S., Luo, M.-L., and 10 others. Antibody against early driver of neurodegeneration cis P-tau blocks brain injury and tauopathy. Nature 523: 431-436, 2015. [PubMed: 26176913] [Full Text: https://doi.org/10.1038/nature14658]

  62. Kwok, J. B. J., Hallupp, M., Loy, C. T., Chan, D. K. Y., Woo, J., Mellick, G. D., Buchanan, D. D., Silburn, P. A., Halliday, G. M., Schofield, P. R. GSK3B polymorphisms alter transcription and splicing in Parkinson's disease. Ann. Neurol. 58: 829-839, 2005. [PubMed: 16315267] [Full Text: https://doi.org/10.1002/ana.20691]

  63. Kwok, J. B. J., Loy, C. T., Hamilton, G., Lau, E., Hallupp, M., Williams, J., Owen, M. J., Broe, G. A., Tang, N., Lam, L., Powell, J. F., Lovestone, S., Schofield, P. R. Glycogen synthase kinase-3 and tau genes interact in Alzheimer's disease. Ann. Neurol. 64: 446-454, 2008. [PubMed: 18991351] [Full Text: https://doi.org/10.1002/ana.21476]

  64. Kwok, J. B. J., Teber, E. T., Loy, C., Hallupp, M., Nicholson, G., Mellick, G. D., Buchanan, D. D., Silburn, P. A., Schofield, P. R. Tau haplotypes regulate transcription and are associated with Parkinson's disease. Ann. Neurol. 55: 329-334, 2004. [PubMed: 14991810] [Full Text: https://doi.org/10.1002/ana.10826]

  65. Lanska, D. J., Currier, R. D., Cohen, M., Gambetti, P., Smith, E. E., Bebin, J., Jackson, J. F., Whitehouse, P. J., Markesbery, W. R. Familial progressive subcortical gliosis. Neurology 44: 1633-1643, 1994. [PubMed: 7936288] [Full Text: https://doi.org/10.1212/wnl.44.9.1633]

  66. Lantos, P. L., Cairns, N. J., Khan, M. N., King, A., Revesz, T., Janssen, J. C., Morris, H., Rossor, M. N. Neuropathologic variation in frontotemporal dementia due to the intronic tau 10 +16 mutation. Neurology 58: 1169-1175, 2002. [PubMed: 11971082] [Full Text: https://doi.org/10.1212/wnl.58.8.1169]

  67. Lei, P., Ayton, S., Finkelstein, D. I., Spoerri, L., Ciccotosto, G. D., Wright, D. K., Wong, B. X. W., Adlard, P. A., Cherny, R. A., Lam, L. Q., Roberts, B. R., Volitakis, I., Egan, G. F., McLean, C. A., Cappai, R., Duce, J. A., Bush, A. I. Tau deficiency induces parkinsonism with dementia by impairing APP-mediated iron export. Nature Med. 18: 291-295, 2012. [PubMed: 22286308] [Full Text: https://doi.org/10.1038/nm.2613]

  68. Lewis, J., Dickson, D. W., Lin, W.-L., Chisholm, L., Corral, A., Jones, G., Yen, S.-H., Sahara, N., Skipper, L., Yager, D., Eckman, C., Hardy, J., Hutton, M., McGowan, E. Enhanced neurofibrillary degeneration in transgenic mice expressing mutant tau and APP. Science 293: 1487-1491, 2001. [PubMed: 11520987] [Full Text: https://doi.org/10.1126/science.1058189]

  69. Lewis, J., McGowan, E., Rockwood, J., Melrose, H., Nacharaju, P., Van Slegtenhorst, M., Gwinn-Hardy, K., Murphy, M. P., Baker, M., Yu, X., Duff, K., Hardy, J., Corral, A., Lin, W.-L., Yen, S.-H., Dickson, D. W., Davies, P., Hutton, M. Neurofibrillary tangles, amyotrophy and progressive motor disturbance in mice expressing mutant (P301L) tau protein. Nature Genet. 25: 402-405, 2000. Note: Erratum: Nature Genet. 26: 127 only, 2000. [PubMed: 10932182] [Full Text: https://doi.org/10.1038/78078]

  70. Li, H.-L., Wang, H.-H., Liu, S.-J., Deng, Y.-Q., Zhang, Y.-J., Tian, Q., Wang, X.-C., Chen, X.-Q., Yang, Y., Zhang, J.-Y., Wang, Q., Xu, H., Liao, F.-F., Wang, J.-Z. Phosphorylation of tau antagonizes apoptosis by stabilizing beta-catenin, a mechanism involved in Alzheimer's neurodegeneration. Proc. Nat. Acad. Sci. 104: 3591-3596, 2007. [PubMed: 17360687] [Full Text: https://doi.org/10.1073/pnas.0609303104]

  71. Lippa, C. F., Zhukareva, V., Kawarai, T., Uryu, K., Shafiq, M., Nee, L. E., Grafman, J., Liang, Y., St George-Hyslop, P. H., Trojanowski, J. Q., Lee, V. M.-Y. Frontotemporal dementia with novel tau pathology and a glu342val tau mutation. Ann. Neurol. 48: 850-858, 2000. [PubMed: 11117541]

  72. Litvan, I., Baker, M., Hutton, M. Tau genotype: no effect on onset, symptom severity, or survival in progressive supranuclear palsy. Neurology 57: 138-140, 2001. [PubMed: 11445645] [Full Text: https://doi.org/10.1212/wnl.57.1.138]

  73. Liu, F., Iqbal, K., Grundke-Iqbal, I., Hart, G. W., Gong, C.-X. O-GlcNAcylation regulates phosphorylation of tau: a mechanism involved in Alzheimer's disease. Proc. Nat. Acad. Sci. 101: 10804-10809, 2004. [PubMed: 15249677] [Full Text: https://doi.org/10.1073/pnas.0400348101]

  74. Lossos, A., Reches, A., Gal, A., Newman, J. P., Soffer, D., Gomori, J. M., Boher, M., Ekstein, D., Biran, I., Meiner, Z., Abramsky, O., Rosenmann, H. Frontotemporal dementia and parkinsonism with the P301S tau gene mutation in a Jewish family. J. Neurol. 250: 733-740, 2003. [PubMed: 12796837] [Full Text: https://doi.org/10.1007/s00415-003-1074-4]

  75. Lu, P.-J., Wulf, G., Zhou, X. Z., Davies, P., Lu, K. P. The prolyl isomerase Pin1 restores the function of Alzheimer-associated phosphorylated tau protein. Nature 399: 784-788, 1999. [PubMed: 10391244] [Full Text: https://doi.org/10.1038/21650]

  76. Lund, H., Cowburn, R. F., Gustafsson, E., Stromberg, K., Svensson, A., Dahllund, L., Malinwosky, D., Sunnemark, D. Tau-tubulin kinase 1 expression, phosphorylation and co-localization with phospho-ser422 tau in Alzheimer's disease brain. Brain Path. 23: 378-389, 2013. [PubMed: 23088643] [Full Text: https://doi.org/10.1111/bpa.12001]

  77. Mamah, C. E., Lesnick, T. G., Lincoln, S. J., Strain, K. J., de Andrade, M., Bower, J. H., Ahlskog, J. E., Rocca, W. A., Farrer, M. J., Maraganore, D. M. Interaction of alpha-synuclein and tau genotypes in Parkinson's disease. Ann. Neurol. 57: 439-443, 2005. [PubMed: 15732111] [Full Text: https://doi.org/10.1002/ana.20387]

  78. Martin, E. R., Scott, W. K., Nance, M. A., Watts, R. L., Hubble, J. P., Koller, W. C., Lyons, K., Pahwa, R., Stern, M. B., Colcher, A., Hiner, B. C., Jankovic, J., and 20 others. Association of single-nucleotide polymorphisms of the Tau gene with late-onset Parkinson disease. JAMA 286: 2245-2250, 2001. [PubMed: 11710889] [Full Text: https://doi.org/10.1001/jama.286.18.2245]

  79. Miyamoto, K., Kowalska, A., Hasegawa, M., Tabira, T., Takahashi, K., Araki, W., Akiguchi, I., Ikemoto, A. Familial frontotemporal dementia and parkinsonism with a novel mutation at an intron 10+11-splice site in the tau gene. Ann. Neurol. 50: 117-120, 2001. [PubMed: 11456301] [Full Text: https://doi.org/10.1002/ana.1083]

  80. Murrell, J. R., Spillantini, M. G., Zolo, P., Guazzelli, M., Smith, M. J., Hasegawa, M., Redi, F., Crowther, R. A., Pietrini, P., Ghetti, B., Goedert, M. Tau gene mutation G389R causes a tauopathy with abundant Pick body-like inclusions and axonal deposits. J. Neuropath. Exp. Neurol. 58: 1207-1226, 1999. [PubMed: 10604746] [Full Text: https://doi.org/10.1097/00005072-199912000-00002]

  81. Myers, A. J., Kaleem, M., Marlowe, L., Pittman, A. M., Lees, A. J., Fung, H. C., Duckworth, J., Leung, D., Gibson, A., Morris, C. M., de Silva, R., Hardy, J. The H1c haplotype at the MAPT locus is associated with Alzheimer's disease. Hum. Molec. Genet. 14: 2399-2404, 2005. [PubMed: 16000317] [Full Text: https://doi.org/10.1093/hmg/ddi241]

  82. Nagase, T., Ishikawa, K., Kikuno, R., Hirosawa, M., Nomura, N., Ohara, O. Prediction of the coding sequences of unidentified human genes. XV. The complete sequences of 100 new cDNA clones from brain which code for large proteins in vitro. DNA Res. 6: 337-345, 1999. [PubMed: 10574462] [Full Text: https://doi.org/10.1093/dnares/6.5.337]

  83. Neumann, M., Schulz-Schaeffer, W., Crowther, R. A., Smith, M. J., Spillantini, M. G., Goedert, M., Kretzschmar, H. A. Pick's disease associated with the novel tau gene mutation K369I. Ann. Neurol. 50: 503-513, 2001. [PubMed: 11601501] [Full Text: https://doi.org/10.1002/ana.1223]

  84. Neve, R. L., Harris, P., Kosik, K. S., Kurnit, D. M., Donlon, T. A. Identification of cDNA clones for the human microtubule-associated protein tau and chromosomal localization of the genes for tau and microtubule-associated protein 2. Brain Res. 387: 271-280, 1986. [PubMed: 3103857] [Full Text: https://doi.org/10.1016/0169-328x(86)90033-1]

  85. Nguyen, M. D., Lariviere, R. C., Julien, J.-P. Deregulation of Cdk5 in a mouse model of ALS: toxicity alleviated by perikaryal neurofilament inclusions. Neuron 30: 135-147, 2001. [PubMed: 11343650] [Full Text: https://doi.org/10.1016/s0896-6273(01)00268-9]

  86. Nicholl, D. J., Greenstone, M. A., Clarke, C. E., Rizzu, P., Crooks, D., Crowe, A., Trojanowski, J. Q., Lee, V. M.-Y., Heutink, P. An English kindred with a novel recessive tauopathy and respiratory failure. Ann. Neurol. 54: 682-686, 2003. [PubMed: 14595660] [Full Text: https://doi.org/10.1002/ana.10747]

  87. Nussbaum, J. M., Schilling, S., Cynis, H., Silva, A., Swanson, E., Wangsanut, T., Tayler, K., Wiltgen, B., Hatami, A., Ronicke, R., Reymann, K., Hutter-Paier, B., Alexandru, A., Jagla, W., Graubner, S., Glabe, C. G., Demuth, H.-U., Bloom, G. S. Prion-like behaviour and tau-dependent cytotoxicity of pyroglutamylated amyloid-beta. Nature 485: 651-655, 2012. [PubMed: 22660329] [Full Text: https://doi.org/10.1038/nature11060]

  88. Oliva, R., Pastor, P. Tau gene delN296 mutation, Parkinson's disease, and atypical supranuclear palsy. (Letter) Ann. Neurol. 55: 448-449, 2004. [PubMed: 14991828] [Full Text: https://doi.org/10.1002/ana.20025]

  89. Panda, D., Samuel, J. C., Massie, M., Feinstein, S. C., Wilson, L. Differential regulation of microtubule dynamics by three- and four-repeat tau: implications for the onset of neurodegenerative disease. Proc. Nat. Acad. Sci. 100: 9548-9553, 2003. [PubMed: 12886013] [Full Text: https://doi.org/10.1073/pnas.1633508100]

  90. Pastor, P., Ezquerra, M., Tolosa, E., Munoz, E., Marti, M. J., Valldeoriola, F., Molinuevo, J. L., Calopa, M., Oliva, R. Further extension of the H1 haplotype associated with progressive supranuclear palsy. Mov. Disord. 17: 550-556, 2002. [PubMed: 12112206] [Full Text: https://doi.org/10.1002/mds.10076]

  91. Pastor, P., Pastor, E., Carnero, C., Vela, R., Garcia, T., Amer, G., Tolosa, E., Oliva, R. Familial atypical progressive supranuclear palsy associated with homozygosity for the delN296 mutation in the tau gene. Ann. Neurol. 49: 263-267, 2001. [PubMed: 11220749] [Full Text: https://doi.org/10.1002/1531-8249(20010201)49:2<263::aid-ana50>3.0.co;2-k]

  92. Petrucelli, L., Dickson, D., Kehoe, K., Taylor, J., Snyder, H., Grover, A., De Lucia, M., McGowan, E., Lewis, J., Prihar, G., Kim, J., Dillmann, W. H., Browne, S. E., Hall, A., Voellmy, R., Tsuboi, Y., Dawson, T. M., Wolozin, B., Hardy, J., Hutton, M. CHIP and Hsp70 regulate tau ubiquitination, degradation and aggregation. Hum. Molec. Genet. 13: 703-714, 2004. [PubMed: 14962978] [Full Text: https://doi.org/10.1093/hmg/ddh083]

  93. Pickering-Brown, S., Baker, M., Bird, T., Trojanowski, J., Lee, V., Morris, H., Rossor, M., Janssen, J. C., Neary, D., Craufurd, D., Richardson, A., Snowden, J., Hardy, J., Mann, D., Hutton, M. Evidence of a founder effect in families with frontotemporal dementia that harbor the tau +16 splice mutation. Am. J. Med. Genet. 125B: 79-82, 2004. [PubMed: 14755449] [Full Text: https://doi.org/10.1002/ajmg.b.20083]

  94. Pickering-Brown, S., Baker, M., Yen, S.-H., Liu, W.-K., Hasegawa, M., Cairns, N., Lantos, P. L., Rossor, M., Iwatsubo, T., Davies, Y., Allsop, D., Furlong, R., Owen, F., Hardy, J., Mann, D., Hutton, M. Pick's disease is associated with mutations in the tau gene. Ann. Neurol. 48: 859-867, 2000. [PubMed: 11117542]

  95. Pickering-Brown, S. M., Richardson, A. M. T., Snowden, J. S., McDonagh, A. M., Burns, A., Braude, W., Baker, M., Liu, W.-K., Yen, S.-H., Hardy, J., Hutton, M., Davies, Y., Allsop, D., Craufurd, D., Neary, D., Mann, D. M. A. Inherited frontotemporal dementia in nine British families associated with intronic mutations in the tau gene. Brain 125: 732-751, 2002. [PubMed: 11912108] [Full Text: https://doi.org/10.1093/brain/awf069]

  96. Pittman, A. M., Myers, A. J., Abou-Sleiman, P., Fung, H. C., Kaleem, M., Marlowe, L., Duckworth, J., Leung, D., Williams, D., Kilford, L., Thomas, N., Morris, C. M., Dickson, D., Wood, N. W., Hardy, J., Lees, A. J., de Silva, R. Linkage disequilibrium fine mapping and haplotype association of the tau gene in progressive supranuclear palsy and corticobasal degeneration. J. Med. Genet. 42: 837-846, 2005. [PubMed: 15792962] [Full Text: https://doi.org/10.1136/jmg.2005.031377]

  97. Pittman, A. M., Myers, A. J., Duckworth, J., Bryden, L., Hanson, M., Abou-Sleiman, P., Wood, N. W., Hardy, J., Lees, A. J., de Silva, R. The structure of the tau haplotype in controls and in progressive supranuclear palsy. Hum. Molec. Genet. 13: 1267-1274, 2004. [PubMed: 15115761] [Full Text: https://doi.org/10.1093/hmg/ddh138]

  98. Poorkaj, P., Bird, T. D., Wijsman, E., Nemens, E., Garruto, R. M., Anderson, L., Andreadis, A., Wiederholt, W. C., Raskind, M., Schellenberg, G. D. Tau is a candidate gene for chromosome 17 frontotemporal dementia. Ann. Neurol. 43: 815-825, 1998. Note: Erratum: Ann. Neurol. 44: 428 only, 1998. [PubMed: 9629852] [Full Text: https://doi.org/10.1002/ana.410430617]

  99. Poorkaj, P., Grossman, M., Steinbart, E., Payami, H., Sadovnick, A., Nochlin, D., Tabira, T., Trojanowski, J. Q., Borson, S., Galasko, D., Reich, S., Quinn, B., Schellenberg, G., Bird, T. D. Frequency of tau gene mutations in familial and sporadic cases of non-Alzheimer dementia. Arch. Neurol. 58: 383-387, 2001. [PubMed: 11255441] [Full Text: https://doi.org/10.1001/archneur.58.3.383]

  100. Poorkaj, P., Kas, A., D'Souza, I., Zhou, Y., Pham, Q., Stone, M., Olson, M. V., Schellenberg, G. D. A genomic sequence analysis of the mouse and human microtubule-associated protein tau. Mammalian Genome 12: 700-712, 2001. [PubMed: 11641718] [Full Text: https://doi.org/10.1007/s00335-001-2044-8]

  101. Poorkaj, P., Muma, N. A., Zhukareva, V., Cochran, E. J., Shannon, K. M., Hurtig, H., Koller, W. C., Bird, T. D., Trojanowski, J. Q., Lee, V. M.-Y., Schellenberg, G. D. An R5L tau mutation in a subject with a progressive supranuclear palsy phenotype. Ann. Neurol. 52: 511-516, 2002. [PubMed: 12325083] [Full Text: https://doi.org/10.1002/ana.10340]

  102. Poorkaj, P. Personal Communication. Seattle, Wash. 11/10/1998.

  103. Przybyla, M., van Eersel, J., van Hummel, A., van der Hoven, J., Sabale, M., Harasta, A., Muller, J., Gajwani, M., Prikas, E., Mueller, T., Stevens, C. H., Power, J., Housley, G. D., Karl, T., Kassiou, M., Ke, Y. D., Ittner, A., Ittner, L. M. Onset of hippocampal network aberration and memory deficits in P301S tau mice are associated with an early gene signature. Brain 143: 1889-1904, 2020. [PubMed: 32375177] [Full Text: https://doi.org/10.1093/brain/awaa133]

  104. Quadros, A., Weeks, O. I., Ait-Ghezala, G. Role of tau in Alzheimer's dementia and other neurodegenerative diseases. J. Appl. Biomed. 5: 1-12, 2007.

  105. Rademakers, R., Cruts, M., van Broeckhoven, C. The role of tau (MAPT) in frontotemporal dementia and related tauopathies. Hum. Mutat. 24: 277-295, 2004. [PubMed: 15365985] [Full Text: https://doi.org/10.1002/humu.20086]

  106. Rademakers, R., Dermaut, B., Peeters, K., Cruts, M., Heutink, P., Goate, A., Van Broeckhoven, C. Tau (MAPT) mutation arg406trp presenting clinically with Alzheimer disease does not share a common founder in western Europe. Hum. Mutat. 22: 409-411, 2003. [PubMed: 14517953] [Full Text: https://doi.org/10.1002/humu.10269]

  107. Rademakers, R., Melquist, S., Cruts, M., Theuns, J., Del-Favero, J., Poorkaj, P., Baker, M., Sleegers, K., Crook, R., De Pooter, T., Bel Kacem, S., Adamson, J., and 15 others. High-density SNP haplotyping suggests altered regulation of tau gene expression in progressive supranuclear palsy. Hum. Molec. Genet. 14: 3281-3292, 2005. [PubMed: 16195395] [Full Text: https://doi.org/10.1093/hmg/ddi361]

  108. Rapoport, M., Dawson, H. N., Binder, L. I., Vitek, M. P., Ferreira, A. Tau is essential to beta-amyloid-induced neurotoxicity. Proc. Nat. Acad. Sci. 99: 6364-6369, 2002. [PubMed: 11959919] [Full Text: https://doi.org/10.1073/pnas.092136199]

  109. Rauch, J. N., Luna, G., Guzman, E., Audouard, M., Challis, C., Sibih, YE., Leshuk, C., Hernandez, I., Wegmann, S., Hyman, B. T., Gradinaru, V., Kampmann, M., Kosik, K. S. LRP1 is a master regulator of tau uptake and spread. Nature 580: 381-385, 2020. [PubMed: 32296178] [Full Text: https://doi.org/10.1038/s41586-020-2156-5]

  110. Reed, L. A., Grabowski, T. J., Schmidt, M. L., Morris, J. C., Goate, A., Solodkin, A., Van Hoesen, G. W., Schelper, R. L., Talbot, C. J., Wragg, M. A., Trojanowski, J. Q. Autosomal dominant dementia with widespread neurofibrillary tangles. Ann. Neurol. 42: 564-572, 1997. [PubMed: 9382467] [Full Text: https://doi.org/10.1002/ana.410420406]

  111. Rizzu, P., Hinkle, D. A., Zhukareva, V., Bonifati, V., Severijnen, L.-A., Martinez, D., Ravid, R., Kamphorst, W., Eberwine, J. H., Lee, V. M.-Y., Trojanowski, J. Q., Heutink, P. DJ-1 colocalizes with tau inclusions: a link between parkinsonism and dementia. Ann. Neurol. 55: 113-118, 2004. [PubMed: 14705119] [Full Text: https://doi.org/10.1002/ana.10782]

  112. Rizzu, P., Van Swieten, J. C., Joosse, M., Hasegawa, M., Stevens, M., Tibben, A., Niermeijer, M. F., Hillebrand, M., Ravid, R., Oostra, B. A., Goedert, M., van Duijn, C. M., Heutink, P. High prevalence of mutations in the microtubule-associated protein tau in a population study of frontotemporal dementia in the Netherlands. Am. J. Hum. Genet. 64: 414-421, 1999. [PubMed: 9973279] [Full Text: https://doi.org/10.1086/302256]

  113. Roberson, E. D., Scearce-Levie, K., Palop, J. J., Yan, F., Cheng, I. H., Wu, T., Gerstein, H., Yu, G.-Q., Mucke, L. Reducing endogenous tau ameliorates amyloid beta-induced deficits in an Alzheimer's disease mouse model. Science 316: 750-754, 2007. [PubMed: 17478722] [Full Text: https://doi.org/10.1126/science.1141736]

  114. Ros, R., Thobois, S., Streichenberger, N., Kopp, N., Sanchez, M. P., Perez, M., Hoenicka, J., Avila, J., Honnorat, J., de Yebenes, J. G. A new mutation of the tau gene, G303V, in early-onset familial progressive supranuclear palsy. Arch. Neurol. 62: 1444-1450, 2005. [PubMed: 16157753] [Full Text: https://doi.org/10.1001/archneur.62.9.1444]

  115. Rossi, G., Gasparoli, E., Pasquali, C., Di Fede, G., Testa, D., Albanese, A., Bracco, F., Tagliavini, F. Progressive supranuclear palsy and Parkinson's disease in a family with a new mutation in the tau gene. (Letter) Ann. Neurol. 55: 448 only, 2004. [PubMed: 14991829] [Full Text: https://doi.org/10.1002/ana.20006]

  116. Rosso, S. M., van Herpen, E., Deelen, W., Kamphorst, W., Severijnen, L.-A., Willemsen, R., Ravid, R., Niermeijer, M. F., Dooijes, D., Smith, M. J., Goedert, M., Heutink, P., van Swieten, J. C. A novel tau mutation, S320F, causes a tauopathy with inclusions similar to those in Pick's disease. Ann. Neurol. 51: 373-376, 2002. [PubMed: 11891833] [Full Text: https://doi.org/10.1002/ana.10140]

  117. Saito, Y., Geyer, A., Sasaki, R., Kuzuhara, S., Nanba, E., Miyasaka, T., Suzuki, K., Murayama, S. Early-onset, rapidly progressive familial tauopathy with R406W mutation. Neurology 58: 811-813, 2002. [PubMed: 11889249] [Full Text: https://doi.org/10.1212/wnl.58.5.811]

  118. SantaCruz, K., Lewis, J., Spires, T., Paulson, J., Kotilinek, L., Ingelsson, M., Guimaraes, A., DeTure, M., Ramsden, M., McGowan, E., Forster, C., Yue, M., Orne, J., Janus, C., Mariash, A., Kuskowski, M., Hyman, B., Hutton, M., Ashe, K. H. Tau suppression in a neurodegenerative mouse model improves memory function. Science 309: 476-481, 2005. [PubMed: 16020737] [Full Text: https://doi.org/10.1126/science.1113694]

  119. Schneider, A., Biernat, J., von Bergen, M., Mandelkow, E., Mandelkow, E. M. Phosphorylation that detaches tau protein from microtubules (Ser262, Ser214) also protects it against aggregation into Alzheimer paired helical filaments. Biochemistry 38: 3549-3558, 1999. [PubMed: 10090741] [Full Text: https://doi.org/10.1021/bi981874p]

  120. Seto-Salvia, N., Clarimon, J., Pagonabarraga, J., Pascual-Sedano, B., Campolongo, A., Combarros, O., Mateo, J. I., Regana, D., Martinez-Corral, M., Marquie, M., Alcolea, D., Suarez-Calvet, M., Molina-Porcel, L., Dols, O., Gomez-Isla, T., Blesa, R., Lleo, A., Kulisevsky, J. Dementia risk in Parkinson disease: disentangling the role of MAPT haplotypes. Arch. Neurol. 68: 359-364, 2011. [PubMed: 21403021] [Full Text: https://doi.org/10.1001/archneurol.2011.17]

  121. Skipper, L., Wilkes, K., Toft, M., Baker, M., Lincoln, S., Hulihan, M., Ross, O. A., Hutton, M., Aasly, J., Farrer, M. Linkage disequilibrium and association of MAPT H1 in Parkinson disease. Am. J. Hum. Genet. 75: 669-677, 2004. [PubMed: 15297935] [Full Text: https://doi.org/10.1086/424492]

  122. Sohn, P. D., Huang, C. T.-L., Yan, R., Fan, L., Tracy, T. E., Camargo, C. M., Montgomery, K. M., Arhar, T., Mok, S.-A., Freilich, R., Baik, J., He, M., Gong, S., Roberson, E. D., Karch, C. M., Gestwicki, J. E., Xu, K., Kosik, K. S., Gan, L. Pathogenic tau impairs axon initial segment plasticity and excitability homeostasis. Neuron 104: 458-470, 2019. [PubMed: 31542321] [Full Text: https://doi.org/10.1016/j.neuron.2019.08.008]

  123. Sperfeld, A. D., Collatz, M. B., Baier, H., Palmbach, M., Storch, A., Schwarz, J., Tatsch, K., Reske, S., Joosse, M., Heutink, P., Ludolph, A. C. FTDP-17: an early-onset phenotype with parkinsonism and epileptic seizures caused by a novel mutation. Ann. Neurol. 46: 708-715, 1999. [PubMed: 10553987] [Full Text: https://doi.org/10.1002/1531-8249(199911)46:5<708::aid-ana5>3.0.co;2-k]

  124. Spillantini, M. G., Murrell, J. R., Goedert, M., Farlow, M. R., Klug, A., Ghetti, B. Mutation in the tau gene in familial multiple system tauopathy with presenile dementia. Proc. Nat. Acad. Sci. 95: 7737-7741, 1998. [PubMed: 9636220] [Full Text: https://doi.org/10.1073/pnas.95.13.7737]

  125. Spillantini, M. G., Yoshida, H., Rizzini, C., Lantos, P. L., Khan, N., Rossor, M. N., Goedert, M., Brown, J. A novel tau mutation (N296N) in familial dementia with swollen achromatic neurons and corticobasal inclusion bodies. Ann. Neurol. 48: 939-943, 2000. [PubMed: 11117553] [Full Text: https://doi.org/10.1002/1531-8249(200012)48:6<939::aid-ana17>3.3.co;2-t]

  126. Spittaels, K., van den Haute, C., van Dorpe, J., Bruynseels, K., Vandezande, K., Laenen, I., Geerts, H., Mercken, M., Sciot, R., van Lommel, A., Loos, R., van Leuven, F. Prominent axonopathy in the brain and spinal cord of transgenic mice overexpressing four-repeat human tau protein. Am. J. Path. 155: 2153-2165, 1999. [PubMed: 10595944] [Full Text: https://doi.org/10.1016/S0002-9440(10)65533-2]

  127. Spittaels, K., van den Haute, C., van Dorpe, J., Geerts, H., Mercken, M., Bruynseels, K., Lasrado, R., Vandezande, K., Laenen, I., Boon, T., van Lint, J., Vandenheede, J, Moechars, D., Loos, R., van Leuven, F. Glycogen synthase kinase-3-beta phosphorylates protein tau and rescues the axonopathy in the central nervous system of human four-repeat tau transgenic mice. J. Biol. Chem. 275: 41340-41349, 2000. [PubMed: 11007782] [Full Text: https://doi.org/10.1074/jbc.M006219200]

  128. Stambolic, V., Ruel, L., Woodgett, J. R. Lithium inhibits glycogen synthase kinase-3 activity and mimics Wingless signalling in intact cells. Curr. Biol. 6: 1664-1668, 1996. Note: Erratum: Curr. Biol. 7: 196 only, 1997. [PubMed: 8994831] [Full Text: https://doi.org/10.1016/s0960-9822(02)70790-2]

  129. Stamer, K., Vogel, R., Thies, E., Mandelkow, E., Mandelkow, E.-M. Tau blocks traffic of organelles, neurofilaments, and APP vesicles in neurons and enhances oxidative stress. J. Cell Biol. 156: 1051-1063, 2002. [PubMed: 11901170] [Full Text: https://doi.org/10.1083/jcb.200108057]

  130. Stefansson, H., Helgason, A., Thorleifsson, G., Steinthorsdottir, V., Masson, G., Barnard, J., Baker, A., Jonasdottir, A., Ingason, A., Gudnadottir, V. G., Desnica, N., Hicks, A., and 15 others. A common inversion under selection in Europeans. Nature Genet. 37: 129-137, 2005. [PubMed: 15654335] [Full Text: https://doi.org/10.1038/ng1508]

  131. Taniguchi, T., Doe, N., Matsuyama, S., Kitamura, Y., Mori, H., Saito, N., Tanaka, C. Transgenic mice expressing mutant (N279K) human tau show mutation dependent cognitive defects without neurofibrillary tangle formation. FEBS Lett. 579: 5704-5712, 2005. [PubMed: 16219306] [Full Text: https://doi.org/10.1016/j.febslet.2005.09.047]

  132. Tatebayashi, Y., Miyasaka, T., Chui, D.-H., Akagi, T., Mishima, K., Iwasaki, K., Fujiwara, M., Tanemura, K., Murayama, M., Ishiguro, K., Planel, E., Sato, S., Hashikawa, T., Takashima, A. Tau filament formation and associative memory deficit in aged mice expressing mutant (R406W) human tau. Proc. Nat. Acad. Sci. 99: 13896-13901, 2002. [PubMed: 12368474] [Full Text: https://doi.org/10.1073/pnas.202205599]

  133. Taylor, L. M., McMillan, P. J., Liachko, N. F., Strovas, T. J., Ghetti, B., Bird, T. D., Keene, C. D., Kraemer, B. C. Pathological phosphorylation of tau and TDP-43 by TTBK1 and TTBK2 drives neurodegeneration. Molec. Neurodegener. 13: 7, 2018. [PubMed: 29409526] [Full Text: https://doi.org/10.1186/s13024-018-0237-9]

  134. Tesseur, I., van Dorpe, J., Bruynseels, K., Bronfman, F., Sciot, R., van Lommel, A., van Leuven, F. Prominent axonopathy and disruption of axonal transport in transgenic mice expressing human apolipoprotein E4 in neurons of brain and spinal cord. Am. J. Path. 157: 1495-1510, 2000. [PubMed: 11073810] [Full Text: https://doi.org/10.1016/S0002-9440(10)64788-8]

  135. Tesseur, I., van Dorpe, J., Spittaels, K., van den Haute, C., Moechars, D., van Leuven, F. Expression of human apolipoprotein E4 in neurons causes hyperphosphorylation of protein tau in the brains of transgenic mice. Am. J. Path. 156: 951-964, 2000. [PubMed: 10702411] [Full Text: https://doi.org/10.1016/S0002-9440(10)64963-2]

  136. Tobin, J. E., Latourelle, J. C., Lew, M. F., Klein, C., Suchowersky, O., Shill, H. A., Golbe, L. I., Mark, M. H., Growdon, J. H., Wooten, G. F., Racette, B. A., Perlmutter, J. S., and 33 others. Haplotypes and gene expression implicate the MAPT region for Parkinson disease: the GenePD Study. Neurology 71: 28-34, 2008. [PubMed: 18509094] [Full Text: https://doi.org/10.1212/01.wnl.0000304051.01650.23]

  137. Tsuboi, Y., Baker, M., Hutton, M. L., Uitti, R. J., Rascol, O., Delisle, M.-B., Soulages, X., Murrell, J. R., Ghetti, B., Yasuda, M., Komure, O., Kuno, S., Arima, K., Sunohara, N., Kobayashi, T., Mizuno, Y., Wszolek, Z. K. Clinical and genetic studies of families with the tau N279K mutation (FTDP-17). Neurology 59: 1791-1793, 2002. [PubMed: 12473774] [Full Text: https://doi.org/10.1212/01.wnl.0000038909.49164.4b]

  138. van Swieten, J. C., Stevens, M., Rosso, S. M., Rizzu, P., Joosse, M., de Koning, I., Kamphorst, W., Ravid, R., Spillantini, M. G., Niermeijer, Heutink, P. Phenotypic variation in hereditary frontotemporal dementia with tau mutations. Ann. Neurol. 46: 617-626, 1999. [PubMed: 10514099] [Full Text: https://doi.org/10.1002/1531-8249(199910)46:4<617::aid-ana10>3.0.co;2-i]

  139. Varani, L., Hasegawa, M., Spillantini, M. G., Smith, M. J., Murrell, J. R., Ghetti, B., Klug, A., Goedert, M., Varani, G. Structure of tau exon 10 splicing regulatory element RNA and destabilization by mutations of frontotemporal dementia and parkinsonism linked to chromosome 17. Proc. Nat. Acad. Sci. 96: 8229-8234, 1999. [PubMed: 10393977] [Full Text: https://doi.org/10.1073/pnas.96.14.8229]

  140. Verpillat, P., Camuzat, A., Hannequin, D., Thomas-Anterion, C., Puel, M., Belliard, S., Dubois, B., Didic, M., Michel, B.-F., Lacomblez, L., Moreaud, O., Sellal, F., Golfier, V., Campion, D., Clerget-Darpoux, F., Brice, A. Association between the extended tau haplotype and frontotemporal dementia. Arch. Neurol. 59: 935-939, 2002. [PubMed: 12056929] [Full Text: https://doi.org/10.1001/archneur.59.6.935]

  141. Volz, A., Boyle, J. M., Cann, H. M., Cottingham, R. W., Orr, H. T., Ziegler, A. Report of the second international workshop in human chromosome 6. Genomics 21: 464-472, 1994. [PubMed: 8088851] [Full Text: https://doi.org/10.1006/geno.1994.1302]

  142. Werber, E., Klein, C., Grunfeld, J., Rabey, J. M. Phenotypic presentation of frontotemporal dementia with parkinsonism-chromosome 17 type P301S in a patient of Jewish-Algerian origin. Mov. Disord. 18: 595-598, 2003. [PubMed: 12722177] [Full Text: https://doi.org/10.1002/mds.10401]

  143. Whitwell, J. L., Jack, C. R., Jr., Boeve, B. F., Senjem, M. L., Baker, M., Ivnik, R. J., Knopman, D. S., Wszolek, Z. K., Petersen, R. C., Rademakers, R., Josephs, K. A. Atrophy patterns in IVS10+16, IVS10+3, N279K, S305N, P301L, and V337M MAPT mutations. Neurology 73: 1058-1065, 2009. [PubMed: 19786698] [Full Text: https://doi.org/10.1212/WNL.0b013e3181b9c8b9]

  144. Wilhelmsen, K. C., Lynch, T., Pavlou, E., Higgins, M., Hygaard, T. G. Localization of disinhibition-dementia-parkinsonism-amyotrophy complex to 17q21-22. Am. J. Hum. Genet. 55: 1159-1165, 1994. [PubMed: 7977375]

  145. Wszolek, Z. K., Pfeiffer, R. F., Bhatt, M. H., Schelper, R. L., Cordes, M., Snow, B. J., Rodnitzky, R. L., Wolters, E. C., Arwert, F., Calne, D. B. Rapidly progressive autosomal dominant parkinsonism and dementia with pallido-ponto-nigral degeneration. Ann. Neurol. 32: 312-320, 1992. [PubMed: 1416801] [Full Text: https://doi.org/10.1002/ana.410320303]

  146. Wszolek, Z. K., Uitti, R. J., Hutton, M. A mutation in the microtubule-associated protein tau in pallido-nigro-luysian degeneration. (Letter) Neurology 54: 2028-2030, 2000. [PubMed: 10822460] [Full Text: https://doi.org/10.1212/wnl.54.10.2028]

  147. Xu, J., Sato, S., Okuyama, S., Swan, R. J., Jacobsen, M. T., Strunk, E., Ikezu, T. Tau-tubulin kinase 1 enhances prefibrillar tau aggregation and motor neuron degeneration in P301L FTDP-17 tau-mutant mice. FASEB J. 24: 2904-2915, 2010. [PubMed: 20354135] [Full Text: https://doi.org/10.1096/fj.09-150144]

  148. Yasuda, M., Kawamata, T., Komure, O., Kuno, S., D'Souza, I., Poorkaj, P., Kawai, J., Tanimukai, S., Yamamoto, Y., Hasegawa, H., Sasahara, M., Hazama, F., Schellenberg, G. D., Tanaka, C. A mutation in the microtubule-associated protein tau in pallido-nigro-luysian degeneration. Neurology 53: 864-868, 1999. [PubMed: 10489057] [Full Text: https://doi.org/10.1212/wnl.53.4.864]

  149. Yasuda, M., Nakamura, Y., Kawamata, T., Kaneyuki, H., Maeda, K., Komure, O. Phenotypic heterogeneity within a new family with the MAPT P301S mutation. Ann. Neurol. 58: 920-928, 2005. [PubMed: 16240366] [Full Text: https://doi.org/10.1002/ana.20668]

  150. Yasuda, M., Yokoyama, K., Nakayasu, T., Nishimura, Y., Matsui, M., Yokoyama, T., Miyoshi, K., Tanaka, C. A Japanese patient with frontotemporal dementia and parkinsonism by a tau P301S mutation. Neurology 55: 1224-1227, 2000. [PubMed: 11071507] [Full Text: https://doi.org/10.1212/wnl.55.8.1224]

  151. Yoshida, H., Crowther, R. A., Goedert, M. Functional effects of tau gene mutations deltaN296 and N296H. J. Neurochem. 80: 548-551, 2002. [PubMed: 11906000] [Full Text: https://doi.org/10.1046/j.0022-3042.2001.00729.x]

  152. Yoshiyama, Y., Higuchi, M., Zhang, B., Huang, S.-M., Iwata, N., Saido, T. C., Maeda, J., Suhara, T., Trojanowski, J. Q., Lee, V. M.-Y. Synapse loss and microglial activation precede tangles in a P301S tauopathy mouse model. Neuron 53: 337-351, 2007. Note: Erratum: Neuron 54: 343-344, 2007. [PubMed: 17270732] [Full Text: https://doi.org/10.1016/j.neuron.2007.01.010]

  153. Zabetian, C. P., Hutter, C. M., Factor, S. A., Nutt, J. G., Higgins, D. S., Griffith, A., Roberts, J. W., Leis, B. C., Kay, D. M., Yearout, D., Montimurro, J. S., Edwards, K. L., Samii, A., Payami, H. Association analysis of MAPT H1 haplotype and subhaplotypes in Parkinson's disease. Ann. Neurol. 62: 137-144, 2007. [PubMed: 17514749] [Full Text: https://doi.org/10.1002/ana.21157]

  154. Zarranz, J. J., Ferrer, I., Lezcano, E., Forcadas, M. I., Eizaguirre, B., Atares, B., Puig, B., Gomez-Esteban, J. C., Fernandez-Maiztegui, C., Rouco, I., Perez-Concha, T., Fernandez, M., Rodriguez, O., Rodriguez-Martinez, A. B., Martinez de Pancorbo, M., Pastor, P., Perez-Tur, J. A novel mutation (K317M) in the MAPT gene causes FTDP and motor neuron disease. Neurology 64: 1578-1585, 2005. [PubMed: 15883319] [Full Text: https://doi.org/10.1212/01.WNL.0000160116.65034.12]

  155. Zhang, W., Tarutani, A., Newell, K. L., Murzin, A. G., Matsubara, T., Falcon, B., Vidal, R., Garringer, H. J., Shi, Y., Ikeuchi, T., Murayama, S., Ghetti, B., Hasegawa, M., Goedert, M., Scheres, S. H. W. Novel tau filament fold in corticobasal degeneration. Nature 580: 283-287, 2020. [PubMed: 32050258] [Full Text: https://doi.org/10.1038/s41586-020-2043-0]

  156. Zhukareva, V., Mann, D., Pickering-Brown, S., Uryu, K., Shuck, T., Shah, K., Grossman, M., Miller, B. L., Hulette, C. M., Feinstein, S. C., Trojanowski, J. Q., Lee, V. M.-Y. Sporadic Pick's disease: a tauopathy characterized by a spectrum of pathological tau isoforms in gray and white matter. Ann. Neurol. 51: 730-739, 2002. [PubMed: 12112079] [Full Text: https://doi.org/10.1002/ana.10222]

  157. Zody, M. C., Jiang, Z., Fung, H.-C., Antonacci, F., Hillier, L. W., Cardone, M. F., Graves, T. A., Kidd, J. M., Cheng, Z., Abouelleil, A., Chen, L., Wallis, J., Glasscock, J., Wilson, R. J., Reily, A. D., Duckworth, J., Ventura, M., Hardy, J., Warren, W. C., Eichler, E. E. Evolutionary toggling of the MAPT 17q21.31 inversion region. Nature Genet. 40: 1076-1083, 2008. [PubMed: 19165922] [Full Text: https://doi.org/10.1038/ng.193]


Contributors:
Bao Lige - updated : 11/09/2023
Bao Lige - updated : 06/30/2021
Bao Lige - updated : 05/05/2021
Ada Hamosh - updated : 08/14/2020
Ada Hamosh - updated : 08/11/2020
Ada Hamosh - updated : 01/06/2020
Ada Hamosh - updated : 09/12/2019
Ada Hamosh - updated : 11/19/2018
George E. Tiller - updated : 09/13/2017
Ada Hamosh - updated : 10/6/2015
George E. Tiller - updated : 9/4/2013
George E. Tiller - updated : 8/14/2013
Ada Hamosh - updated : 7/19/2012
Cassandra L. Kniffin - updated : 3/6/2012
Cassandra L. Kniffin - updated : 11/14/2011
Cassandra L. Kniffin - updated : 3/23/2011
Patricia A. Hartz - updated : 2/3/2011
Ada Hamosh - updated : 5/26/2010
Cassandra L. Kniffin - updated : 3/24/2010
George E. Tiller - updated : 2/22/2010
Cassandra L. Kniffin - updated : 1/4/2010
Cassandra L. Kniffin - updated : 12/14/2009
George E. Tiller - updated : 10/23/2009
Cassandra L. Kniffin - updated : 3/16/2009
George E. Tiller - updated : 1/12/2009
Ada Hamosh - updated : 10/30/2008
Cassandra L. Kniffin - updated : 10/17/2008
Cassandra L. Kniffin - updated : 4/18/2008
Ada Hamosh - updated : 4/4/2008
Cassandra L. Kniffin - updated : 1/7/2008
Cassandra L. Kniffin - updated : 10/11/2007
Ada Hamosh - updated : 5/30/2007
Victor A. McKusick - updated : 3/29/2007
Cassandra L. Kniffin - updated : 3/15/2007
George E. Tiller - updated : 10/5/2006
George E. Tiller - updated : 9/7/2006
Cassandra L. Kniffin - updated : 4/20/2006
Cassandra L. Kniffin - updated : 3/13/2006
Cassandra L. Kniffin - updated : 3/6/2006
Cassandra L. Kniffin -updated : 11/3/2005
Anne M. Stumpf - updated : 9/27/2005
Cassandra L. Kniffin - updated : 9/1/2005
Ada Hamosh - updated : 8/15/2005
Cassandra L. Kniffin - updated : 7/19/2005
Cassandra L. Kniffin - updated : 1/20/2005
Victor A. McKusick - updated : 12/9/2004
Cassandra L. Kniffin - updated : 10/27/2004
Victor A. McKusick - updated : 9/21/2004
Victor A. McKusick - updated : 9/9/2004
Cassandra L. Kniffin - reorganized : 6/10/2004
Cassandra L. Kniffin - updated : 1/7/2004
Victor A. McKusick - updated : 11/19/2003
Victor A. McKusick - updated : 9/8/2003
Cassandra L. Kniffin - updated : 4/29/2003
Cassandra L. Kniffin - updated : 3/6/2003
Cassandra L. Kniffin - updated : 2/11/2003
Cassandra L. Kniffin - updated : 1/21/2003
Victor A. McKusick - updated : 1/8/2003
Cassandra L. Kniffin - updated : 12/6/2002
Cassandra L. Kniffin - updated : 11/26/2002
Victor A. McKusick - updated : 11/22/2002
Dawn Watkins-Chow - updated : 11/5/2002
Cassandra L. Kniffin - updated : 10/2/2002
Dawn Watkins-Chow - updated : 8/22/2002
Cassandra L. Kniffin - updated : 7/29/2002
Victor A. McKusick - updated : 6/17/2002
Victor A. McKusick - updated : 6/6/2002
Cassandra L. Kniffin - updated : 6/5/2002
Victor A. McKusick - updated : 6/4/2002
Victor A. McKusick - updated : 12/4/2001
Dawn Watkins-Chow - updated : 11/25/2001
Ada Hamosh - updated : 9/12/2001
Victor A. McKusick - updated : 9/28/2000
Victor A. McKusick - updated : 7/31/2000
George E. Tiller - updated : 5/2/2000
Victor A. McKusick - updated : 3/24/2000
Ada Hamosh - updated : 2/22/2000
Victor A. McKusick - updated : 8/10/1999
Ada Hamosh - updated : 6/24/1999
Victor A. McKusick - updated : 4/7/1999
Victor A. McKusick - updated : 2/26/1999
Victor A. McKusick - updated : 2/18/1999
Victor A. McKusick - updated : 12/3/1998
Victor A. McKusick - updated : 11/18/1998
Victor A. McKusick - updated : 11/2/1998
Victor A. McKusick - updated : 8/20/1998
Victor A. McKusick - updated : 6/29/1998
Moyra Smith - updated : 7/1/1996
Andre K. Cheng - edited : 4/23/1996
Orest Hurko - updated : 4/4/1996

Creation Date:
Victor A. McKusick : 4/15/1987

Edit History:
mgross : 11/09/2023
mgross : 06/30/2021
carol : 05/06/2021
mgross : 05/05/2021
alopez : 08/14/2020
alopez : 08/11/2020
carol : 08/11/2020
carol : 01/07/2020
alopez : 01/06/2020
alopez : 09/12/2019
alopez : 11/19/2018
alopez : 09/13/2017
carol : 08/18/2016
joanna : 03/24/2016
carol : 1/11/2016
alopez : 10/6/2015
mgross : 9/17/2015
carol : 4/13/2015
alopez : 1/26/2015
alopez : 9/10/2013
tpirozzi : 9/4/2013
tpirozzi : 8/16/2013
tpirozzi : 8/16/2013
tpirozzi : 8/15/2013
tpirozzi : 8/15/2013
tpirozzi : 8/14/2013
terry : 4/4/2013
terry : 3/14/2013
terry : 11/29/2012
alopez : 7/23/2012
terry : 7/19/2012
carol : 6/8/2012
carol : 5/16/2012
carol : 3/9/2012
ckniffin : 3/6/2012
carol : 11/16/2011
terry : 11/16/2011
ckniffin : 11/14/2011
carol : 4/18/2011
carol : 4/14/2011
wwang : 4/12/2011
ckniffin : 3/23/2011
mgross : 2/9/2011
terry : 2/3/2011
wwang : 1/7/2011
ckniffin : 12/10/2010
alopez : 12/8/2010
ckniffin : 11/17/2010
alopez : 6/1/2010
terry : 5/26/2010
carol : 3/24/2010
ckniffin : 3/24/2010
wwang : 3/2/2010
terry : 2/22/2010
wwang : 1/20/2010
ckniffin : 1/4/2010
wwang : 12/28/2009
carol : 12/23/2009
ckniffin : 12/15/2009
ckniffin : 12/14/2009
wwang : 11/3/2009
terry : 10/23/2009
wwang : 8/24/2009
terry : 8/12/2009
terry : 6/3/2009
alopez : 4/15/2009
wwang : 3/26/2009
ckniffin : 3/16/2009
wwang : 1/12/2009
alopez : 10/30/2008
wwang : 10/27/2008
ckniffin : 10/17/2008
wwang : 4/23/2008
ckniffin : 4/18/2008
alopez : 4/11/2008
terry : 4/4/2008
carol : 2/29/2008
wwang : 1/23/2008
ckniffin : 1/7/2008
wwang : 10/19/2007
ckniffin : 10/11/2007
alopez : 5/30/2007
terry : 5/30/2007
terry : 5/9/2007
carol : 3/29/2007
carol : 3/29/2007
ckniffin : 3/15/2007
alopez : 10/5/2006
alopez : 9/7/2006
carol : 6/21/2006
ckniffin : 6/12/2006
wwang : 4/25/2006
ckniffin : 4/20/2006
alopez : 4/12/2006
wwang : 3/20/2006
ckniffin : 3/13/2006
wwang : 3/13/2006
ckniffin : 3/6/2006
alopez : 2/3/2006
wwang : 11/11/2005
ckniffin : 11/3/2005
alopez : 9/27/2005
wwang : 9/6/2005
ckniffin : 9/1/2005
carol : 8/16/2005
terry : 8/15/2005
wwang : 7/26/2005
ckniffin : 7/19/2005
terry : 7/11/2005
terry : 2/22/2005
tkritzer : 1/25/2005
ckniffin : 1/20/2005
tkritzer : 1/6/2005
terry : 12/9/2004
terry : 11/2/2004
tkritzer : 11/1/2004
ckniffin : 10/27/2004
tkritzer : 9/23/2004
terry : 9/21/2004
tkritzer : 9/10/2004
terry : 9/9/2004
carol : 6/10/2004
ckniffin : 6/9/2004
ckniffin : 6/8/2004
mgross : 3/17/2004
carol : 1/29/2004
ckniffin : 1/7/2004
tkritzer : 12/19/2003
tkritzer : 12/18/2003
tkritzer : 11/24/2003
terry : 11/19/2003
cwells : 9/10/2003
terry : 9/8/2003
tkritzer : 5/21/2003
tkritzer : 4/29/2003
ckniffin : 4/29/2003
carol : 3/17/2003
tkritzer : 3/14/2003
ckniffin : 3/6/2003
carol : 2/25/2003
carol : 2/25/2003
ckniffin : 2/11/2003
carol : 1/31/2003
tkritzer : 1/28/2003
ckniffin : 1/21/2003
cwells : 1/9/2003
tkritzer : 1/8/2003
carol : 12/6/2002
ckniffin : 12/6/2002
cwells : 12/3/2002
ckniffin : 11/26/2002
cwells : 11/22/2002
terry : 11/20/2002
carol : 11/7/2002
tkritzer : 11/5/2002
tkritzer : 11/5/2002
carol : 10/31/2002
tkritzer : 10/25/2002
carol : 10/21/2002
ckniffin : 10/4/2002
ckniffin : 10/4/2002
ckniffin : 10/2/2002
tkritzer : 8/22/2002
carol : 8/7/2002
ckniffin : 8/7/2002
ckniffin : 7/29/2002
mgross : 6/25/2002
terry : 6/21/2002
terry : 6/17/2002
alopez : 6/12/2002
mgross : 6/10/2002
mgross : 6/10/2002
terry : 6/6/2002
carol : 6/5/2002
ckniffin : 6/5/2002
ckniffin : 6/5/2002
terry : 6/4/2002
carol : 12/10/2001
mcapotos : 12/4/2001
carol : 11/25/2001
alopez : 9/13/2001
alopez : 9/13/2001
terry : 9/12/2001
alopez : 10/30/2000
mcapotos : 10/17/2000
mcapotos : 10/16/2000
terry : 10/6/2000
terry : 9/28/2000
alopez : 8/31/2000
alopez : 7/31/2000
terry : 7/31/2000
alopez : 5/2/2000
mcapotos : 4/18/2000
mcapotos : 4/14/2000
terry : 3/24/2000
alopez : 2/22/2000
alopez : 8/23/1999
terry : 8/10/1999
alopez : 8/5/1999
alopez : 6/24/1999
alopez : 6/24/1999
carol : 4/7/1999
terry : 2/26/1999
carol : 2/22/1999
terry : 2/18/1999
alopez : 12/4/1998
alopez : 12/3/1998
terry : 11/18/1998
carol : 11/11/1998
carol : 11/10/1998
terry : 11/2/1998
terry : 8/20/1998
dholmes : 7/22/1998
carol : 6/29/1998
terry : 6/29/1998
mark : 9/19/1996
mark : 7/1/1996
terry : 7/1/1996
terry : 7/1/1996
mark : 4/23/1996
mark : 4/23/1996
terry : 4/15/1996
mark : 4/4/1996
carol : 9/22/1993
supermim : 3/16/1992
supermim : 3/20/1990
ddp : 10/27/1989
root : 6/27/1988
root : 6/20/1988