Abstract
INTRODUCTION
Alternative splicing of the human MAPT gene generates six brain‐specific TAU isoforms. Imbalances in the TAU isoform ratio can lead to neurodegenerative diseases, underscoring the need for precise control over TAU isoform balance. Tauopathies, characterized by intracellular aggregates of hyperphosphorylated TAU, exhibit extensive neurodegeneration and can be classified by the TAU isoforms present in pathological accumulations.
METHODS
A comprehensive review of TAU and related dementia syndromes literature was conducted using PubMed, Google Scholar, and preprint server.
RESULTS
While TAU is recognized as key driver of neurodegeneration in specific tauopathies, the contribution of the isoforms to neuronal function and disease development remains largely elusive.
DISCUSSION
In this review we describe the role of TAU isoforms in health and disease, and stress the importance of comprehending and studying TAU isoforms in both, physiological and pathological context, in order to develop targeted therapeutic interventions for TAU‐associated diseases.
Highlights
MAPT splicing is tightly regulated during neuronal maturation and throughout life.
TAU isoform expression is development‐, cell‐type and brain region specific.
The contribution of TAU to neurodegeneration might be isoform‐specific.
Ineffective TAU‐based therapies highlight the need for specific targeting strategies.
Keywords: alternative splicing, Alzheimer's disease, genetic tauopathy, MAPT, TAU isoforms, tauopathy
1. INTRODUCTION
The microtubule‐associated protein TAU plays a pivotal role in the regulation of fundamental neuronal processes. TAU's manifold physiological roles include stabilizing microtubules (MTs), facilitating axonal transport, and modulating synaptic plasticity. 1 Under physiological conditions TAU supports neuronal plasticity by regulating the structural dynamics of MTs necessary for neuronal growth and function. The complexity of TAU functions is further enhanced by alternative splicing of the TAU‐encoding MAPT gene, theoretically enabling the expression of over 30 different splice isoforms in the human central and peripheral nervous system, each of which could potentially exhibit distinct functions. 2 , 3 , 4 , 5 Notably, MAPT alternative splicing and TAU isoform expression are highly species dependent. While six isoforms are expressed in the adult human brain (2N4R, 1N4R, 0N4R, 2N3R, 1N3R, 0N3R), only three are detected in adult rodents (2N4R, 1N4R, 0N4R) (Table 1). 3 , 6 , 7 , 8
TABLE 1.
TAU isoform expression in human and murine brains.
| Parameter | Human TAU | Murine TAU | ||||||
|---|---|---|---|---|---|---|---|---|
| Fetal isoform | 0N3R | 0N3R | ||||||
| Adult isoforms (CNS) | Length [aa] | Molecular weight [kDa] | Expression level * | Length [aa] | Molecular weight [kDa] | Expression level* | ||
| 0N3R | 352 | 36.76 | 10%–20 % | |||||
| 0N4R | 383 | 40.01 | 8%–13% | 0N4R | 350 | 36.74 | 20% | |
| 1N3R | 381 | 39.72 | 23%–30% | |||||
| 1N4R | 412 | 42.97 | 20%–25% | 1N4R | 372 | 38.96 | 40% | |
| 2N3R | 410 | 42.60 | 3%–5% | |||||
| 2N4R | 441 | 45.85 | 5%–15% | 2N4R | 430 | 44.89 | 40% | |
The importance of TAU exceeds its physiological role, as dysregulation and loss of the microtubule‐stabilizing function has been intricately linked to neurodegenerative diseases (NDDs) (reviewed, e.g., in 9 , 10 ). In sporadic forms of Alzheimer's disease (AD), TAU undergoes abnormal phosphorylation (often referred to as hyperphosphorylation) and eventually changes its subcellular localization, leading to its accumulation and the formation of neurofibrillary tangles (NFTs). 11 , 12 , 13 , 14 , 15 , 16 TAU pathology has been observed in a variety of neurodegenerative diseases, such as frontotemporal dementia (FTD), progressive supranuclear palsy (PSP), Pick's disease (PiD), and corticobasal degeneration (CBD), therefore termed tauopathies. The majority of these diseases occur sporadically, and the underlying triggers remain elusive. 17 However, over 50 variants in MAPT have been associated with FTD (Alzforum database 12/2023), which i.a. lead to imbalances in TAU isoform expression 18 , 19 , 20 , 21 (reviewed, e.g., in 22 , 23 ), clearly demonstrating a pathogenic role of TAU isoforms in at least a subset of FTD. Tauopathies can be further classified according to the TAU isoforms present in the pathological inclusions. For instance, PiD is characterized by tangles containing 3R‐TAU isoforms (0N3R, 1N3R, and 2N3R), whereas 4R‐TAU (0N4R, 1N4R, and 2N4R) accumulates in disorders like PSP and CBD. In AD, aggregates consist of all TAU isoforms (as reviewed in 24 , 25 ). Despite the pivotal role of TAU in the pathogenesis of these disorders, our understanding of the contribution of individual TAU isoforms remains limited. Most studies have predominantly focused on “full‐length” or mature 2N4R‐TAU or specific mutants associated with FTD, and often rely on rodent model systems that express only 4R‐TAU. While whole classification systems of tauopathies are based on subsets of isoforms present in the tauopathy‐typical accumulations of TAU (e.g., 3R and 4R tauopathies), the role the isoforms in physiology and disease is underappreciated. We here point out that TAU isoforms are not only important for the established histopathological classification, but apparently also for physiological neuronal and brain development, intracellular distribution, and mediation of neuronal dysfunction. We should appreciate the differences of the human TAU isoforms for the design of preclinical studies (in particular for the use and generation of transgenic mice, which are currently based overwhelmingly on single and not representative isoforms), which may well pave the way for the selective targeting of a subset of TAU species (e.g. isoforms) as a therapeutic avenue.
RESEARCH IN CONTEXT
Systematic review: The authors reviewed the literature using databases, such as PubMed and Google Scholar, meeting abstracts, and preprint servers with a special focus on brain TAU isoforms, TAU biology, and TAU‐related neurodegenerative dementia syndromes.
Interpretation: While the TAU protein is intricately linked to pathophysiology of several tauopathies, the relevance and contribution of its isoforms remains relatively unexplored. Our review summarizes the current state of TAU research with a special focus on the human brain‐specific TAU isoforms, their physiological characteristics, and their implications in tauopathies and neurodegenerative diseases.
Future directions: The failure of a variety of TAU‐targeting therapies in tauopathies highlights the urgent need for improved targeting strategies. Our review emphasizes the importance of comprehending and studying TAU isoforms in both physiological and pathological context in order to develop targeted therapeutic interventions for TAU‐related dementia syndromes that address TAU pathology while maintaining essential cellular TAU functions.
Hence, in this review, we summarize the diverse physiological functions of the six major brain TAU isoforms, the implications of alternative splicing, and variations observed across species. Furthermore, we describe the pathological role of TAU isoforms in NDDs, with a special focus on AD, and stress the importance of comprehending and studying TAU isoforms in both physiological and pathological context in order to develop targeted therapeutic interventions for TAU‐associated NDDs.
2. MAPT ALTERNATIVE SPLICING AND TAU ISOFORMS
The TAU protein is a natively unfolded, intrinsically disordered, soluble protein, which is abundantly expressed in the central and peripheral nervous system, especially in neurons. However low levels of TAU can be also detected in astrocytes and oligodendrocytes. 26 , 27 , 28 , 29 , 30 , 31 , 32 In addition, TAU is widely expressed in other tissues, such as the pancreas, breast, kidney, and skeletal muscle, however at much lower levels than in neuronal tissues. 33 , 34
The human TAU protein is encoded by the MAPT gene on chromosome 17q21.31. Alternative splicing gives rise to a diverse array of isoforms in the human central and peripheral nervous system, with over 30 different predicted isoforms. 2 , 3 , 4 , 5 The primary targets of the alternative splicing process are exons 2, 3, and 10, resulting in the expression of six brain‐associated TAU isoforms in mature human neurons characterized by the number of N‐terminal inserts (0N, 1N, or 2N) and C‐terminal repeat domains (3R or 4R) (Figure 1 and Table 1). 2 Exon 10 inclusion or exclusion determines the number of C‐terminal repeat domains, while splicing of exons 2 and 3 determine the number of N‐terminal inserts. Of note, exon 3 inclusion depends on the inclusion of exon 2 due to a weak branch point that favors exclusion of exon 3 from the final mRNA transcript. 33 , 35
FIGURE 1.

Alternative splicing of the human MAPT gene. The human MAPT gene encodes for six brain‐specific TAU isoforms based on alternative splicing of exons 2, 3, and 10. The TAU protein can be structured into four different domains: The N‐terminal projection domain, which projects away from the microtubules (MTs), acts as a spacer and interacts with components of the plasma membrane. The proline‐rich region is involved in cellular signalling by interacting with Src‐familiy kinases such as Fyn. The MT‐binding domain mediates MT polymerization and stability, and the C‐terminal region additionally contributes to MT polymerization and TAU's interaction with the plasma membrane. Figure created with BioRender.com
While the shortest isoform, 0N3R, is exclusively expressed during neurogenesis, the isoform composition switches upon neuronal maturation toward larger TAU isoforms, particularly 2N (comprising 2N3R and 2N4R) and 4R (comprising 0N4R, 1N4R, and 2N4R), resulting in nearly equal ratios of 3R and 4R‐TAU isoform expression in adult brains. 8 , 36 Even though 3R and 4R isoforms are present in nearly equal amounts, the relative abundance of 0N, 1N, and 2N TAU isoforms in humans differs significantly: While 1N TAU isoforms account for approx. 50% of all TAU species, 0 and 2N TAU isoforms each make up only up to 20% (Table 1). 37 Further, the isoform composition and protein abundance can differ in brain regions and cellular subtypes. 37 , 38 , 39 , 40 Notably, inclusion of exon 4a upon neuronal maturation results in the additional expression of a high molecular weight TAU isoform in the peripheral nervous system, referred to as big TAU, which might play a role in the regulation of axonal transport of large projecting neurons. 5 , 6 , 41 , 42
MAPT is not exclusively expressed in the brain, but TAU can be found in a variety of tissues. Briefly, several studies indeed show that TAU found in the periphery, in particular the peripheral nervous system, so‐called‚ “Big TAU,” includes exon 4a (5), but other isoforms are expressed as well. Interestingly, the fetal brain isoform 0N3R is usually not expressed elsewhere, but data are limited to adult tissue. 43 , 44 One study specifically looking at the enteric nervous system (ENS) detected expression of brain‐typical TAU isoforms (but not big TAU), indicating that the ENS may have particular, brainlike properties. In rats, the adult ENS did not express 0N3R‐TAU, while primary cultures thereof (which usually maintain some degree of immaturity) did, indicating that, at least in some peripheral tissues, also the brain‐typical isoforms are expressed. 45 Also, in Western blots of typical PNS‐tissues (like rat sciatic nerve), there are usually faint bands at the size of the brain‐specific isoforms, suggesting that the inclusion of the typical PNS‐exons is far from perfect. 45 We note that exon 4 is expressed in the PNS, exon 6 is expressed in the brain and the spinal cord, while for exon 8, there are no data that would allow to deduce its expression in a specific or any tissue. 33 Expression of MAPT in non‐brain tissues and the PNS has received very little interest, so data are scarce (see recent reviews 41 , 42 ), and often unclear. Of note, expression of TAU in cell lines derived from pancreatic beta cells have been detected 46 , 47 , but not in the actual tissue. As pancreatic beta cells share important features with neurons, it is conceivable that in paradigms of de‐differentiation, creation of cell lines or cancerous behavior, these cells may express TAU, but not under physiological conditions.
What are molecular factors mediating TAU splicing? Over 10 serine/arginine‐rich splicing factors (SRs) and heterogenous nuclear ribonucleoproteins (hnRNPs) are implicated in the splicing of exons 2, 3, and 10, 48 , 49 , 50 and developmental stage specific splicing inhibitors have been described in the regulation of exon 10 splicing. 50 , 51 The ratio of 3R to 4R TAU isoforms is crucial for normal neuronal function and imbalances are directly linked to various neurodegenerative diseases (see also section 3. The contribution of TAU isoforms to tauopathies). Notably, exon 10 is a focal point of splicing regulation due to its pivotal role in familial tauopathies like progressive supranuclear palsy (PSP), corticobasal degeneration (CBD), Pick's disease (PiD), and frontotemporal dementia and parkinsonism linked to chromosome 17 (FTDP‐17). 48 , 49 , 52 Pathogenic variants at the exon 10‐intron 10 boundary disturb proper exon 10 splicing leading to an altered 3R and 4R TAU isoform balance and are directly associated with FTD. 18 , 19 , 20 , 48 , 49 , 53 , 54 , 55 , 56 Under normal conditions, exon 10 splicing is regulated by a variety of cis‐ and trans‐acting elements. Deletion of the SC35‐like element at the 5′ end of exon 10, for example, promotes exon 10 inclusion in several cell culture models. 57 Furthermore a polypurine enhancer (PPE) and an A/C rich enhancer (ACE) further add to the complex regulation of exon 10 splicing regulation, and mutations in these elements are directly linked to FTD and other neurodegenerative diseases by promoting exon 10 inclusion. 58 , 59 , 60 Due to high self‐complementary between the 3′ end of exon 10 and the 5′ end of intron 10, a stem loop is usually formed that increases availability for U1 snRNP and enhancing 4R TAU expression. 61 , 62 Exon 10 alternative splicing is further regulated by various SR proteins, that either promote (e.g., SRSF1, SRSF2, SC35), SRSR6, and SRSR9, or suppress its inclusion (e.g., SRSF3, SRSF4, SRSF11) (reviewed, e.g., in 33 ). Kinases regulating the activity of SR proteins, such as Dyrk1A, 57 , 63 PKA, 64 and CLK1, 2, and 3 65 have been additionally shown to regulate the alternative splicing of MAPT. Further proteins affecting exon 10 splicing are the polypyrimidine tract‐binding Protein 1 (PTBP1) that represses exon 10 inclusion, 66 and Nova‐1. 67 Muscleblind‐like (MBNL) proteins are involved in developmental splicing and could potentially influence MAPT splicing, although direct evidence remains sparse. 68 , 69 Their known roles in modulating splicing in various contexts make them candidates for future research in TAU isoform regulation. The same is true for CELF2 (also known as ETR3): CELF2 belongs to the CELF family, known for its involvement in developmental splicing regulation. While its direct role in TAU splicing is not extensively documented, CELF proteins’ broad regulatory functions suggest a possible influence on MAPT splicing patterns. 68
MicroRNAs (miRNAs) regulate MAPT alternative splicing by targeting specific sites within the mRNA. Several brain miRNAs, including miR‐124, miR‐9, miR‐132, and miR‐137, have been identified as important regulators of MAPT alternative splicing, particularly in the context of exon 10 splicing (reviewed, e.g., in 70 ). These miRNAs can indirectly regulate the expression of exon 10 trans‐acting factors, thereby affecting the balance between 3R and 4R TAU isoforms. While the specific mechanisms by which these miRNAs regulate MAPT alternative splicing are still being elucidated, it is clear that miRNAs play a significant role in modulating the splicing of TAU pre‐mRNA, particularly in the context of tauopathies (reviewed in 71 , 72 ).
For the technical and molecular mechanism of MAPT alternative splicing, we refer the interested reader to recent exhaustive reviews (for example 33 , 50 , 68 , 70 , 71 , 72 , 73 , 74 , 75 ).
In addition to the above mentioned splicing factors, two major haplotypes influence MAPT gene expression and alternative splicing. 76 The H1 haplotype, found in approximately 75% of the European population, promotes exon 10 inclusion and increases susceptibility to tauopathies like Parkinson's disease (PD), PSP, and CBD. 76 , 77 , 78 , 79 , 80 , 81 Conversely, the H2 haplotype, characterized by a ∼970 kb inversion spanning the entire MAPT locus, reduces AD risk and enhances exon 3 inclusion. 82 , 83 , 84 The H2 haplotype is nearly exclusive to the European population, with an occurrence rate of around 25%. 85
Although the coding sequence of MAPT is relatively conserved among mammals, and humans and rodents share the same intron‐exon organization, the splicing pattern diverges significantly across phylogenetic groups 3 , 6 , 38 , 86 (reviewed, e.g., in 7 ). In adult human and primate brains, six TAU isoforms (0N3R, 0N4R, 1N3R, 1N4R, 2N3R, and 2N4R) are present, whereas in murine brains only up to four TAU isoforms (0N3R, 0N4R, 1N4R, 2N4R) are expressed (Figure 1, Table 1). Similar to human neuronal development, murine neuronal development is characterized by exclusive expression of 0N3R; however, adult murine neurons predominantly contain 4R TAU isoforms (0N4R, 1N4R, and 2N4R) with roughly equal expression ratios (Table 1). 6 , 38 , 86 , 87 , 88 These dissimilarities may at least partially explain human vulnerability to MAPT splicing abnormalities and the absence of TAU pathology in many transgenic mouse models of AD, as summarized, for example, in 25 , 89 .
In summary, the intricate splicing patterns and diverse isoform landscape of the TAU protein underscore its central role in both physiological neuronal function and the pathogenesis of neurodegenerative diseases. The differential expression of TAU isoforms in different brain regions, cell types, and species underscores the complexity of its regulation and suggests specific vulnerabilities in certain contexts.
3. STRUCTURE, LOCALIZATION, AND PRIMARY FUNCTION OF THE TAU ISOFORMS
3.1. TAU domain structure and canonical primary function of microtubule stabilization
The TAU protein is structured into four different domains, depending on the amino acid sequence and the interactions with other proteins 90 , 91 : (1) the N‐terminal projection domain comprising the N‐terminal inserts and is hence different for 0N, 1N, and 2N isoforms, (2) a proline‐rich region consistent among CNS isoforms, (3) the MT‐binding domain that harbors either three (3R) or four (4R) repeat domains, and (4) the C‐terminal region, which is identical between the CNS isoforms (Figure 1). The six human brain‐specific TAU isoforms differ in their size ranging from 352 aa for the smallest, fetal TAU isoform 0N3R, to 441 aa for the largest 2N4R‐TAU isoform (Table 1). While 3R‐TAU isoforms comprise three MT‐binding repeats (R1, R3, and R4), 4R‐TAU isoforms have an additional repeat domain (R2), due to the inclusion of exon 10 during alternative splicing. Under physiological conditions, TAU is mainly bound to neuronal MTs, however, TAU phosphorylation at specific sites (in particular the KxGS‐motifs phosphorylated by the microtubule affinity regulating kinases (MARKs) (see, e.g., 92 ), but also hyperphosphorylation under pathological conditions lead to detachment from MTs. Underhyperphosphorylated conditions, a “paperclip” fold can be observed, in particular for the largest TAU isoform (2N4R). This fold is based on interactions between the N‐terminus with the repeat domains (R1‐R4) and increases the aggregation propensity of this TAU species. 93 , 94 , 95 Together with the proline‐rich and the C‐terminal pseudo‐repeat region, the MT‐binding domain enables tubulin binding and regulates MT‐polymerization efficiency, while the N‐terminal region projects away from MTs, which could regulate MT‐spacing. 96 The differences in the N‐terminal exons may thus result in different distances between MTs, but this has not been addressed experimentally.
By binding to monomeric tubulin subunits, TAU promotes the assembly and stability of MTs, also enabling them to have long labile domains necessary for rapid MT assembly and disassembly, 97 without evidence of differential impact of the different isoforms. Complete loss of TAU (e.g., in Mapt KO mice or MAPT KO in human neurons) does not lead to overt loss of MTs, but only to subtle defects in neuronal outgrowth (which is usually connected to MT function). 98 , 99 , 100 , 101 It is thus obvious that loss of TAU does not automatically cause loss or destabilization of MTs. The concept that TAU facilitates dynamic lability of MTs in physiological conditions (see, e.g., 97 , 102 ), and the concept that TAU (as well as i.a. many other axonal MAPs and MAP‐associated kinases) certainly play a role not only in MT stability, but also in MT dynamics, underlines the experimental evidence that TAU may be crucial for proper MT dynamics in physiological conditions. Accordingly, loss of TAU might not directly lead to MT destabilization, but rather towards a disturbed ratio between labile and stable MT domains. Either way, even if TAU makes MTs more dynamic (or more stable), pathological TAU (which does not necessarily mean loss of TAU) may result in “overdynamic” or fragile MT, which may even result in excessive severing of MTs, as outlined previously, 101 , 103 which could cause MT loss. TAU further serves as a spacer and anchors MTs to the cellular plasma membrane via its interaction with actin and annexins. 104 , 105 , 106
MT‐binding affinity is tightly regulated by phosphorylation, with 85 predicted phosphorylatable sites (reviewed in 107 ), and more than 50 of which have already been demonstrated experimentally. 108 , 109 , 110 Most phosphorylation sites identified in post mortem brains are located within the isoform independent proline‐rich and C‐terminal region. 110 While the N‐terminal domain is also often phosphorylated, there has been so far no experimental evidence of phosphorylation in the second insert of TAU, which may be due to the lower abundance of 2N‐isoforms, but could also hint toward joint regulation of 1N/2N isoforms versus 0N isoforms, as there are many phosphorylated sites experimentally confirmed in the first insert (e.g., S46, S56, S68, T69, T71). In contrast, there are few phosphorylation sites present in the repeat domains, and those present (in particular the KxGS‐motifs) may be redundant. The second MT‐binding repeat domain (only present in 4R‐TAU) contains three phosphorylation sites (S289, S293, S305) phosphorylated in post mortem brains. Due to the repetitive nature of the repeat domains, however, specific or individual regulation seems unlikely (for summary and discussion about PTMs of TAU, see 111 ). The additional MT‐binding repeat in 4R‐TAU isoforms does increase their MT‐binding affinity, leading to faster and more efficient MT assembly compared to 3R‐TAU isoforms in vitro. 112 , 113 , 114 These differential effects, however, have not been observed in human neuroblastoma cells (SH‐SY5Y) expressing individual TAU isoforms. 115 In this model, 4R‐TAU isoforms reduced the cellular size, whereas 3R‐TAU isoforms resulted in an increase, but no significant differences in regard to MT number, stability, length, and growth rate were observed between individual isoforms. 115 This suggests that all isoforms can mediate the canonical function of TAU, i.e. MT stabilization, at least in an isolated context in human neuron‐like cells.
3.2. TAU sorting and localization
In the developing brain and particularly during neuronal polarization, TAU is efficiently sorted into the axon. Nevertheless, in the mature brain, a minor proportion is also detectable in the soma, dendrites, and was even reported in the nucleus. 116 , 117 , 118 , 119 , 120 , 121 , 122 Axonal enrichment of TAU is likely achieved by various strategies, such as active transport of MAPT mRNA mediated by an axonal localization signal in the 3′ untranslated region (UTR), free protein diffusion, active protein transport along MTs, rapid degradation of somatic TAU, and inhibition of retrograde diffusion by a TAU diffusion barrier (TDB) at the axon initial segment (AIS) 123 , 124 , 125 , 126 , 127 , 128 , 129 , 130 ; for review see, e.g., 131 . Furthermore, the MT‐binding affinity, which is influenced by posttranslational modifications (PTMs) like phosphorylation and acetylation, tends to be higher in the axon, and dependent on the number of C‐terminal repeat domains encoded by exon 10. 74 , 112 , 113 , 114 , 132 , 133 , 134 Compartment‐specific enzymes, e.g. the p38γ kinase at the postsynaptic density 135 or PP2A phosphatase in dendrites, 136 , 137 enable differential modification of TAU and fine‐tuning of its MT‐affinity and subcellular localization. Overall, TAU is considered an axonal protein and is often used as a marker of such. However, several studies have demonstrated that the axonal sorting efficiency among the six TAU isoforms differs significantly: While the smaller human 0N isoforms are efficiently sorted into axons, the 1N and 2N isoforms tend to be partially retained in dendrites and cell bodies of rodent primary and hiPSC‐derived neurons. 100 , 124 , 138 In the murine brain, 0N4R and 2N4R TAU isoforms are enriched in the CA3 mossy fiber region that contains axonal projections from dentate granule cells. 139 Further, 1N4R and 2N4R isoforms are enriched in the dendritic fractions, which could be explained by the stronger interaction of these isoforms with dendritic or postsynaptic proteins. 139 TAU phosphorylation at isoform‐independent epitopes, such as T231 (AT180), S262 (12E8), and S396/404 (PHF1) were shown to regulate dendritic localization of TAU. 140 Furthermore, a truncated TAU variant lacking the MT‐binding domain is uniformly distributed in the neuron. 140 Therefore, the differences in the subcellular localization of the TAU isoforms likely depend on the presence and number of N‐terminal inserts, which could be explained by differences in PTMs, MT‐binding affinity, and protein‐protein interactions of TAU isoforms (see also the TAU isoform‐specific functions beyond MT stabilization section). 117 , 132 , 133 , 134 Further, the isoform composition and protein abundance can differ in brain regions and cellular subtypes, 37 , 38 , 39 , 40 indicative of tight brain region‐, cellular compartment‐, and neuronal subtype‐specific regulation.
4. TAU ISOFORM‐SPECIFIC FUNCTIONS BEYOND MT STABILIZATION
In addition to its primary function in stabilizing neuronal MTs, TAU interacts with various other proteins, influencing a wide range of cellular processes (Figure 2A) (for review see 25 , 141 , 142 ). Limited data are available on the functional diversity of individual TAU isoforms. Several studies have revealed distinct intracellular distributions of TAU isoforms in primary neuronal cultures, mouse brains, and hiPSC‐derived neurons, indicating their potential differential roles in TAU‐related functions (see the TAU sorting and localization section). 100 , 124 , 139 In accordance with these differential localizations, co‐immunoprecipitation (Co‐IP) experiments using endogenous murine 0N4R, 1N4R, and 2N4R‐TAU isoforms have revealed distinct isoform‐specific interactions: Proteins binding to murine 0N4R‐TAU were enriched in cellular homeostasis, the cellular respiration pathway, and glycolysis pathway and include α‐ and β‐synuclein, synapsin‐1, and synaptogyrin‐3. 1N4R‐TAU binding proteins, such as ATPase, neuromodulin, calmodulin, and sodium‐ and chloride‐dependent GABA transporter 3, were also enriched in glycolysis, while proteins that preferentially bind to 2N4R‐TAU were involved in ATP synthesis and synaptic transmission, including, for example, ApoA1, ApoE, and synaptotagmin. 143 , 144 However, specific interaction domains or motifs have not been identified that could explain the differential binding of TAU isoforms to these proteins. Notably, also none such isoform‐specific interactome studies have been conducted in human brain lysates or neurons to date.
FIGURE 2.

TAU functions in health and disease. In healthy neurons, TAU is enriched in the axon, where it binds and enables MTs to form long labile/ dynamic domains. A small fraction of TAU is reported to be localized in in the soma, dendrites and the nucleus. Interactome and functional studies highlight the diverse roles of TAU in various cellular contexts. For example, by interacting with DNA and RNA, TAU can influence transcription and translation. 142 Dendritic TAU enhances dendrite and spine maturation and may be important for synaptic activity. During neuronal development, TAU facilitates axonal outgrowth and growth cone dynamics and later has a role in synapse formation and synaptic transmission. Under pathological conditions, such as extracellular aggregation of amyloid beta plaques, TAU dissociates from axonal MTs, accumulates in in the somatodendritic compartment, and gets hyperphosphorylated. Due to the phosphorylation, TAU changes its conformation, which increases its aggregation propensity and results in the accumulation of TAU assemblies in NFTs. Reduction of the MT‐bound TAU results in changes of MT dynamics and leads to impairments of downstream functions, such as axonal transport, which ultimately drives synaptic dysfunction and neuronal cell death. Furthermore, TAU missorting into the dendrites causes postsynaptic spine loss and NMDA receptor mediated excitotoxicity driving cognitive decline. Furthermore, somatic or missorted TAU can inhibit translation, and cause proteasomal and mitochondrial dysfunction. At the pre‐synapse TAU further contributes to synaptic dysfunction and can spread trans‐synaptically to receiving neurons, possibly mediating the progression of the pathology. Figure created with BioRender.com
A recent study using hiPSC‐derived neurons revealed that TAU interactions encompass a broad spectrum of functions beyond cytoskeletal organization, and include pre‐synaptic vesicle dynamics, proteasomal processes, RNA binding, and mitochondrial activities, implying diverse additional roles for TAU (Figure 2A). 144 Interestingly, differences between the interactome of N‐ and C‐terminally APEX‐tagged 2N4R‐TAU were observed: N‐terminally tagged TAU interacted mainly with active zone proteins, such as Dynamin 1, α‐/β‐SNAP, RAB3GAP1, and Liprin α3, and with postsynaptic proteins like ABI2, Cadherin‐2, and Nectin, while C‐terminally tagged TAU was involved in vesicle fusion and interacted with proteins, such as Syntaxin 1A/1B, RAB3A, RIMS1, and Mint1. Furthermore, N‐terminally tagged TAU interacted with a SNARE protein present in autophagosomes, YKT6, which plays important roles in lysosomal fusion. 144 Differential labeling of TAU interacting proteins by N‐ and C‐ terminal APEX point toward subdomain‐, and likely isoform‐specific interactions, which might depend on the number of N‐terminal inserts and C‐terminal repeat domains that potentially are influenced by changes in the alternative splicing pattern of the MAPT gene. However, these likely domain‐specific interactions have not been further addressed in this study. In SH‐SY5Y cells, overexpressed 2N4R‐TAU interactions where significantly enriched for components of the translational and RNA‐processing machinery. 145 Using truncated TAU variants, the study further unraveled that RNA‐binding proteins (RNPs), ribosomal subunits, and heteronuclear ribonucleoproteins (hnRNPs) among others, preferentially bind to the N‐terminal domains (in this case containing both inserts (exon 2 and exon 3), while 14‐3‐3 proteins, heat‐shock proteins (HSPs), and actin‐related proteins interact mainly with the C‐terminal domains of 2N4R‐TAU. 145 A study using IMR neuroblastoma cells and human neurons differentiated from a neural progenitor cell line (ReN VM), that both express equal amounts of 2N4R and 2N3R, further consistently showed the interactions of TAU with HSPs, 14‐3‐3 proteins, hnRNPs, and RNPs. 146 The studies performed in human cellular context are in line with results obtained from human post‐mortem tissue and consistently identified proteasomal proteins and RNPs, such as hnRNPs, FUS, SFPQ, and PTBP1, to interact with TAU. 142 , 147 , 148 , 149 In a comprehensive study using the HT22 murine neuronal cell line cells, 2N4R, 1N4R, and 0N4R TAU demonstrated different capabilities to undergo liquid‐liquid phase separation (LLPS), 150 a process used for membrane‐less compartmentalization, and implicated in pathological protein aggregation observed in neurodegenerative diseases (reviewed in 151 ). While p62 positive condensates were formed upon presence of 2N4R TAU, these were absent upon expression of 1N4R and 0N4R, suggesting that these isoforms are involved in other physiological processes. 150
Although human and murine TAU are highly similar due to strong MAPT sequence homology, 3 , 6 , 7 , 25 the interactome of TAU differs remarkably between species (reviewed in 142 ), which hints towards differential functions of TAU between these two species. In a rodent context, a variety of TAU species have been studied, ranging from WT rat TAU (which corresponds to all 4R‐TAU isoforms) to transgenic mice overexpressing human 1N4R‐ or 2N4R‐TAU with FTD‐associated mutations (P301S or P301L). Likely due to these differences, both in isoforms and pathogenicity, the TAU interactome varies strongly between individual studies. Nevertheless, interactions of TAU with HSPs, and synaptic and trafficking components are consistently found, suggesting further roles of TAU in these processes. 142 , 152 , 153 , 154 , 155 Notably, a significant amount of the interacting partners identified in human studies did not appear in the screenings performed in rodent models, however some interactions of TAU with vesicle proteins and proteins involved in the unfolded protein response remain consistent between human and rodents (reviewed in 142 ).
The vast majority of functions and interactions of TAU have been primarily studied in pathological contexts, either using TAU variants associated with FTD, or by using cellular stressors, such as amyloid beta oligomers (see also the The contribution of TAU isoforms to AD pathology section), but with no attention to the individual isoforms. Briefly, in these studies, TAU has been directly linked to impairments of the proteasomal system, reduced protein synthesis by interacting with ribosomal components, excitotoxicity by interaction with components of the NMDA receptor complex and spine loss by destabilization of local MTs (Figure 2B). 101 , 142 , 147 , 156 , 157 , 158 , 159 , 160 , 161 , 162 , 163 , 164 Furthermore, nuclear localization of TAU has been observed. While TAU protects RNA and DNA under physiological conditions, the interaction of TAU with RNPs and RNA can also increase its aggregation propensity. 88 , 165 , 166 , 167 , 168 , 169 , 170 TAU further plays a role in regulating synaptic plasticity at the postsynaptic site through its interactions with NMDA and AMPA receptors, as well as other components of the postsynaptic density, including PSD95, Fyn kinase, and GSK‐3β. 156 , 159 , 164 , 171 , 172 Here, the role of the individual isoforms remains unresolved.
In summary, the functions of TAU extend beyond MT stabilization to encompass diverse roles in cellular processes. Isoform‐specific studies in murine brains together with data from human neurons reveal unique patterns of interactions, emphasizing the importance of considering different TAU isoforms when studying TAU physiology. Recent investigations in hiPSC‐derived neurons further highlight the involvement of TAU in pre‐synaptic vesicle dynamics, proteasomal processes, translation, and mitochondrial function, expanding the functional spectrum. Despite high sequence homology, the TAU interactome differs between species, emphasizing the need for species‐specific considerations. While pathological implications have been extensively studied, the roles of TAU isoforms in physiological conditions, such as translational regulation and synaptic plasticity, remain relatively unexplored. Collectively, these findings advance our understanding of the diverse contributions of TAU to cellular function, with potential implications for health and disease.
5. THE EFFECT OF TAU DEPLETION AND (RE‐)EXPRESSION OF (INDIVIDUAL) TAU ISOFORMS
Through MT‐binding, TAU influences various neuronal processes such as axonal differentiation, morphogenesis, outgrowth, branching, cargo transport, and neuronal plasticity. 97 , 103 , 119 , 173 , 174 , 175 TAU depletion is in general well tolerated in rodents, possibly due to the compensatory upregulation of other microtubule‐associated proteins (MAPs), such as MAP1A and MAP6 during neuronal development. 97 , 99 , 102 , 176 However, Mapt knockout (KO) mice show some strain‐dependent mild age‐related characteristics, such as changes in sleep‐wake patterns, motor deficits (parkinsonism), deterioration of cardiovascular function, and impairments of fear conditioning. 177 , 178 , 179 , 180 , 181 , 182 , 183 , 184
In contrast, TAU depletion in adult animals enhances hippocampal neurogenesis and leads to resistance against depressive behaviour. 185 In primary cultures, KO, knockdown (KD), or antibody mediated suppression of TAU however results in a mild but detectable reduction of axonal growth, decreased MT density, reduced neurite length, and reduced synaptic spines in some studies, but does not cause axonal transport deficits, changes in MT‐stability, neuronal activity and synaptic spines. 97 , 99 , 101 , 176 , 186 , 187 , 188 Similar effects have also been observed in hiPSC‐derived neurons, which in addition exhibit changes in the AIS but do not show impairments in neuronal activity. 100 , 189 , 190 The expression of all six human TAU isoforms from an artificial chromosome in murine Mapt KO background resulted in increased dendritic localization of TAU, TAU hyperphosphorylation and aggregation, memory deficits, and neuronal loss. 191 , 192 , 193 , 194 These phenotypes are likely caused by the imbalance between 3R and 4R expression observed in these animals, which is also associated with several tauopathies 191 , 192 , 193 , 194 (see the TAU isoforms in tauopathies section). In contrast, humanization of the entire murine Mapt locus similarly resulted in a human‐like expression of TAU isoforms but does not lead to any overt phenotype or impairments in Mapt KO animals. 195 Heterozygous deletion of exon 10 (encoding for the fourth repeat domain) in mice resulted in a one‐to‐one ratio of 3R and 4R murine TAU isoform expression, mimicking the physiological abundance of TAU in human brains. While the deletion of exon 10 did not result in apparent morphological changes of the brain, an age‐dependent decline of sensorimotor function was observed in these animals, 196 which hint toward an isoform‐specific haploinsufficiency pathomechanism. While this is in line with a recent study claiming MAPT‐haploinsufficiency‐based loss of dopaminergic neurons and decreased viability due to a lack of MAP1A compensation, 197 there is to date little evidence supporting (isoform‐specific) TAU haploinsufficiency. There is, however, a clear selection against loss of function (LoF) mutations in the brain‐development‐specific isoform 0N3R‐TAU (but not the other isoforms), indicating that in humans, and at least during neuronal development, haploinsufficiency of 0N3R‐TAU is not tolerated or prevents further procreation (e.g., because of severe enough intellectual disability). 198 However, these results highlight the importance of a tight regulation of TAU isoform expression, underline the differences between human and murine Mapt splicing and the importance of a (species‐specific) splicing pattern for proper physiological function of TAU.
The effects of individual TAU isoforms are also remarkably different depending on the model system studied, the technique used to express TAU and which isoform is chosen. Expression of human 2N4R‐TAU in a murine Mapt KO background increased hippocampal neurogenesis and improved cognitive function compared to TAU depleted animals. 199 , 200 Notably, mild toxic effects of overexpressed human 0N4R‐TAU have been observed, which induced Golgi fragmentation in isolated murine hippocampal neurons. 201 Furthermore, overexpression of human TAU isoforms in WT primary murine neurons significantly affected spine and dendrite maturation: especially 2N4R‐TAU expression enhanced neuronal maturation, accelerated spine formation and dendritic growth. 124 Depletion of TAU in the SH‐SY5Y neuroblastoma cell line, resulted in an increased resistance to induced DNA double‐strand breaks (DSBs) but lead to cellular senescence in a p53 dependent manner. 202 The re‐expression of 2N4R‐ or 2N3R‐TAU, restored cellular sensitivity to DSBs and ameliorated the pro‐survival effects of TAU depletion, but other isoforms were not evaluated. 202 Re‐expression of individual TAU isoforms in hiPSC‐derived MAPT KO neurons restored neurite length and AIS length to wild‐type (WT) levels and did not affect neuronal activity. 100
All in all, the results from human TAU knock‐in (KI) and murine Mapt KO animals and derived primary neurons, and recent evidence from hiPSC‐derived MAPT‐KO neurons, suggest that TAU depletion can be tolerated relatively well in rodents likely because of compensation by other MAPs, however, also implies important modulatory roles of TAU in neuronal development and maturation, in particular in human cells. In humans, mutations in MAPT which lead to a loss of its primary function by altering the MT binding affinity or reduce its ability to promote MT assembly, are causative for FTD, but usually also come along with a gain of function, namely increased aggregation propensity. 198 , 203 Furthermore, tauopathies can be classified according to the TAU isoforms present in neuronal inclusions (see also the TAU isoforms in other tauopathies section). In humans, true haploinsufficiency of MAPT alone has not been reported. Microdeletions at chromosome 17q21.3, which encompassed MAPT and CRHR1 (corticotropin‐releasing hormone receptor 1, a gene that is currently not associated with a genetic disease), 204 , 205 were later found to also include the gene KANSL1, which is causative for the phenotype of the (KANSL1‐associated) Koolen‐De Vries syndrome, independent of the absence or presence of MAPT. 206 This, and the fact that up to date there are no homozygous LoF mutation carriers of MAPT reported, hint towards at least some selection against (haplo‐) insufficiency or LoF mutations in human development, but whether LoF is a contributor to (age‐associated) neurodegeneration is unclear. This underlines the differences between human and murine neurons regarding the tolerance of TAU loss and the importance of TAU function especially during neuronal development.
6. THE CONTRIBUTION OF TAU ISOFORMS TO TAUOPATHIES
Tauopathies are a heterogenous group of more than 20 neurodegenerative diseases characterized by mislocalization and accumulation of hyperphosphorylated TAU in the somatodendritic compartment. 7 They are categorized based on the presence or absence of co‐pathologies, the genetic background, the TAU isoforms present in the insoluble aggregates, or the conformation of the TAU filaments (reviewed in 9 , 25 , 207 , 208 , 209 ). Primary tauopathies, such as PSP, CBD, PiD, and FTDP‐17, are diseases in which TAU pathology is the major contributing factor or hallmark. In contrast, secondary tauopathies, such as AD, PD, and Huntington's disease (HD) involve the aggregation of an additional amyloidogenic protein (Amyloid beta (Aβ), α‐synuclein, or huntingtin, respectively). The majority of tauopathies are sporadic, and endo‐lysosomal and mitochondrial dysfunction, changes in splice factors, aberrant ROS production, and epigenetic dysregulations together with certain risk factors, such as TAU haplotype, an unhealthy lifestyle (e.g., alcohol abuse, smoking, overweight), cardiovascular factors (such as hypertension or diabetes) and head injuries have been implicated in the disease pathogenesis (reviewed in 207 , 210 , 211 , 212 ).
Over 50 MAPT variants are linked to FTD spectrumdisorders, contributing to up to 20% of familial cases (Alzforum database 12/2023) (reviewed in 22 , 25 , 209 ). These mutations, often centered around exons 9‐12, or found within intronic regions, frequently lead to impaired splicing of exon 10. This disruption results in an imbalance between 3R and 4R TAU isoforms (e.g., ΔK280, ΔN296, P301L, P301S). 54 , 213 Other missense mutations or deletions diminish the MT binding affinity of TAU, reduce its ability to promote MT assembly (e.g., P301L, P301S, V337M), enhance protein aggregation propensity (e.g., ΔK280 and P301L), disrupt interactions with other proteins (e.g., R406W), or alter its phosphorylation pattern (e.g., P301L, V337M, G272V). 21 , 25 , 213 , 214 , 215 Moreover, the dysfunction of other proteins has been associated with the misregulation of TAU splicing in sporadic tauopathies, which leads to an elevated inclusion of exon 10, and resulting in a higher expression of 4R TAU isoforms. 216 , 217 Depending on the altered splicing pattern, tauopathies can be classified as either 3R or 4R tauopathies. For example, PiD is distinguished by tangles containing 3R isoforms (0N3R, 1N3R, and 2N3R), while 4R‐TAU (0N4R, 1N4R, and 2N4R) accumulates in disorders such as PSP and CBD. In AD patients, aggregates comprise all isoforms. 24 , 25 The disease‐specific aggregation of TAU isoforms can be explained by differences in the seeding potential of individual isoforms: 3R TAU seeds, for example, can recruit both 3R and 4R monomers, whereas 4R isoform seeds contain only 4R monomers. 218 , 219 , 220 Of note, TAU aggregation propensity is significantly affected by its PTMs and the specific isoform(s) expressed (see the The contribution of TAU isoforms to AD pathology section). A recent in vitro study demonstrated, for example, that while acetylation can specifically drive 3R‐TAU aggregation, acetylation of lysine residues within the second MT‐binding repeat unique to 4R‐TAU isoforms prevents TAU fibril formation. 221 Cryo‐EM characterization of fibrils derived from brains of patients with distinct tauopathies revealed that distinct TAU filament structures characterize different tauopathies, implying substantial variations in the mechanisms underlying these diseases and underscoring the necessity for disease‐modifying treatments that account for these variations. 208 , 222 , 223 , 224 , 225
In myotonic dystrophy type 1 (DM1), a CTG microsatellite repeat expansion in the 3′ UTR of dystrophia myotonica protein kinase (DMPK) gene causes the expression of a toxic RNA. This RNA aggregates as so‐called nuclear foci and ultimately results in aberrant splicing of various genes, leading to multisystemic symptoms including heterogeneous brain involvement. 226 , 227 Notably, NFTs of DM1 patients exhibit a selective buildup of the 0N3R TAU isoform due to altered splicing of the MAPT gene, primarily marked by reduced exon 2 and 3 inclusion. 228 , 229 , 230 Dysregulation of the splicing factor embryonic lethal abnormal vision‐like RNA‐binding protein‐3 (ETR3/CELF2) has been shown to repress exon 2, 3, and 10 inclusion in DM1, thereby contributing to the pathological TAU protein accumulation and NFT formation. 227 , 229 , 231 , 232 In addition, MBNL1 and MBNL2 dysregulation disrupts the developmental regulation of TAU splicing, leading to missplicing of TAU exon 2. 228 , 229 , 233
The contribution of the TAU isoforms to pathological processes has been mainly studied in the context of FTD by the use of mutated TAU variants, which are directly associated with genetic forms of the disease. In human neurons generated from an immortalized neural progenitor cell line (ReN VM), the presence of mutant 2N4R‐TAUP301L in addition to endogenously expressed WT 3R‐TAU (mainly 0N3R) led to the disruption of specific TAU interactions with non‐muscle myosins. 146 These myosins play crucial roles in regulating the morphology of dendritic spines, underlining the essential role of TAU in maintaining dendritic spine integrity. 146 , 234 , 235 , 236 Additionally, in hiPSC‐derived neurons, the presence of 2N4R‐TAUP301L or 2N4R‐TAUV337M resulted in a reduction of TAU interactions with ribosomal and mitochondrial proteins, causing impairments in mitochondrial biogenesis. 144 Moreover, a decrease in the levels of these disrupted TAU‐interacting proteins observed in vitro was also observed in AD patient brains and correlated with the progression of AD. 144 Expression of 2N4RP301L in SH‐SY5Y neuroblastoma cells resulted in a reduced interaction of TAU and proteasomal subunits but increased interactions with ribosomal subunits, translation initiation factors and hnRNPs. 145 Comparison of the WT rat 2N4R‐TAU interactome with a truncated human 4R‐TAU (hTAU151‐391) variant in transgenic rats lead to the identification of novel interacting proteins, for example, Baiap2, Gpr37l1, and Nptx1, that were further validated by Co‐IP of human AD brains and could play a role in TAU pathology. 152 Enhanced expression of 0N4R‐ and 1N4R‐TAU in hiPSC‐derived neurons mediated by introducing point mutations at the 3′ and 5′ ends of exon 10 that prevent splicing of the pre‐mRNA, did not result in obvious phenotypes and treatment with aggregation‐prone TAU seeds did not induce TAU accumulation. 237 However, when neurons express 0N4RP301S and 1N4R‐TAUP301S under the same conditions, progressive TAU accumulation can be observed. Expression of mutant TAU further resulted in impairments of endo‐lysosomal function and neuronal activity. In addition, a recent CRISPRi screen in these neurons identified components of the endo‐lysosomal system and proteins mediating UFMylation of TAU as novel factors promoting the initial seeding and propagation of TAU. 237
Proteomic analysis of murine neurons expressing human 0N4RP301L revealed changes in the interaction of TAU with RNPs, that upon disease progression specifically co‐localize with TAU inclusions. 153 Further, 1N4RP301L expression in mice confirmed interactions of TAU with HSPs and found a novel link between the ubiquitin‐proteasomal system and TAU pathology. 155 In a Drosophila melanogaster tauopathy model, overexpression of the 0N4R isoform led to neurodegeneration and impairments in learning and memory. 238
In conclusion, studies in diverse cellular models, including hiPSC‐derived neurons and transgenic rodents, highlight isoform‐specific interactions, and their relevance to disease progression. It is crucial to note that the studies discussed here predominantly utilized isolated and mutated TAU isoforms for their analyses to uncover novel disease mechanisms. While some findings are consistent between studies, the question of whether these mechanisms are truly isoform‐specific remains elusive, especially concerning that often isolated 2N4R or 1N4R TAU isoforms are studied. The considerable differences between studies, including variations in the isoform studied, the use of FTD‐associated mutants, and the diverse model systems employed, underscore the complexity of TAU in disease pathology. The relevance of these findings to disease pathology and the functional role of TAU in this context therefore remains to be explored.
7. AD PATHOLOGY: GENERAL INTRODUCTION AND ISOFORM‐SPECIFIC CONTRIBUTIONS OF TAU
In general, AD is a slowly progressive disease that is characterized by the decline of memory, language, visuospatial, and executive functions over time. 239 Although individuals in the initial phases remain cognitively normal, slight changes in the brain can be detected many years to decades before symptoms manifest. 11 , 13 , 239 In the advanced stages of AD, individuals lose the ability to perform daily tasks independently, often requiring full‐time care. At the molecular level, the following key characteristics of AD (as reviewed elsewhere, 11 , 12 , 13 , 240 ) can be detected in the brain: (1) The deposition of amyloid beta (Aβ) peptides in insoluble extracellular amyloid plaques and intracellular aggregation of hyperphosphorylated TAU protein in NFTs and neuropil threads. (2) Significant synaptic loss and neuronal cell death which correlate with the onset of cognitive impairment and severity of symptoms. (3) Impairment of the blood‐brain barrier integrity, resulting in the activation of microglia and astrocytes and subsequent neuroinflammation.
7.1. Familial and sporadic AD: Risk factors and their impact on amyloid pathology
Briefly, the majority of AD cases emerge in a sporadic fashion, primarily affecting individuals over the age of 65 (late‐onset /sporadic AD). Nevertheless, a smaller proportion of cases (< 5%), results from autosomal dominant inheritance (familial AD), associated with early onset of the disease between the age of 30 and 65. Pathogenic variants are identified in the genes encoding for the amyloid precursor protein (APP) or proteases involved in APP processing (PSEN1 or PSEN2). These variations lead to an enhanced production of aggregation prone Aβ peptides, or directly influence the propensity for Aβ aggregation 241 (reviewed, e.g., in 11 , 242 ). These peptides, namely Aβ40 and Aβ42, which accumulate in insoluble extracellular plaques in AD, are produced through sequential cleavage of APP by the β‐secretase and γ‐secretase complex. 243 , 244 Under normal physiological conditions and at low concentrations, Aβ peptides facilitate synapse formation and synaptic signaling. The processing and release of Aβ peptides is therefore closely linked to synaptic activity. 245 , 246 , 247 , 248 , 249 , 250 At higher concentrations, Aβ peptides, particularly Aβ42, may aggregate into β‐sheet conformations such as oligomers and (proto) fibrils leading to significant downstream consequences, including neuronal dysfunction (mainly through interaction with NMDA and insulin receptors), TAU hyperphosphorylation, and synapse loss, that ultimately result in neurodegeneration 247 (reviewed, e.g., in 251 , 252 , 253 ). Furthermore, various APP mutations are linked to familial AD, for instance, the Swedish (K670_M671delinsNL), London (V717I), or Indiana (V717F) mutations that result in amplified production of amyloidogenic Aβ peptides or structural modifications which enhance the propensity of the peptides to aggregate (reviewed in 243 ). Additionally, variations in PSEN1 or PSEN2, which encode for presenilin 1 (PS1) and 2 (PS2) that are part of the γ‐secretase complex, result in a dominant‐negative effect on γ‐secretase activity and alter the balance between Aβ42 and Aβ40 peptides. 254 , 255
More than 40 alleles have been implied as susceptibility factors for late‐onset sporadic AD (reviewed in 11 ). The APOE ɛ4 allele, the most prominent risk allele, for example, increases the AD risk up to 15 times when present on both alleles. 256 Other risk genes, such as TREM2, SORL1, ABCA7, BIN1, CD33, and CLU, are primarily associated with APP processing, immune response, microglial function, cholesterol lipid dysfunction, endocytosis, and vascular factors, implying central roles of these pathways for disease pathology. 257 , 258 , 259 , 260 Interestingly, variants in MAPT are not associated with an increased risk for AD, although several of TAU's interaction partners, such as BIN1, FERMT2, and PTK2B, have been identified as risk factors. 259
On the other hand, several protective alleles that reduce the risk of developing AD, have been identified, including the APOE ɛ2 allele leading to a two times reduced risk of AD. 261 , 262 In addition, the rare Icelandic APP mutation (A647T) has been shown to protect against cognitive decline by reducing the production of Aβ42 by up to 40%. 263 Remarkably, a homozygous mutation in APOE ɛ3 (R136S, “Christchurch”), identified in one individual with an inherited PSEN1 variant, protected against neurodegeneration and cognitive decline for decades after the expected onset of disease symptoms. 264
7.2. The general role of TAU as a disease mediator in AD
While TAU protein pathology has been intricately linked to AD and, in contrast to Aβ pathology, correlates well with the cognitive decline observed in patients, no disease‐causing variants in MAPT have been associated with AD and co‐occurring Aβ pathology is necessary to drive TAU accumulation and disease progression. 265 , 266 , 267 , 268 , 269 The significance of TAU in disease progression is underscored by the absence of cognitive symptoms in some individuals with Aβ pathology and the protection of Mapt KO animals from Aβ‐induced neurotoxicity, memory impairments, and premature mortality. 98 , 270 , 271 , 272 , 273 , 274 , 275 , 276 , 277 TAU reduction in hAPP, APP23, and pR5 transgenic mice for example prevented memory loss, premature mortality, and reduced excitotoxicity induced by Aβ. 164 , 270 This is likely mediated by the reduced expression of the tyrosine kinase Fyn upon TAU depletion, as Fyn KO animals exhibit the same protective effect as Mapt KO mice. 278 This protective mechanism can be observed also in primary cultures and slice cultures from TAU depleted animals, which show sustained long‐term potentiation (LTP) and functional axonal transport of mitochondria and vesicles. 164 , 279 This effect has been recently also demonstrated in MAPT KO hiPSC‐derived neurons that are protected from Aβ oligomer (AβO)‐induced neuronal toxicity. 100 , 190 Of note, the protective effect of TAU reduction or depletion is specific to AD and does not protect from pathology in animal models of other tauopathies, such as ALS and PD, and even exacerbates the phenotype of Niemann‐Pick disease type C models. 279 , 280 , 281
In contrast to Aβ peptides, TAU pathology manifests as intracellular NFTs, neuropil threads, and dystrophic neurites. The pattern of TAU pathology in the brain of AD patients is highly reproducible and follows determined trajectories: TAU accumulation appears first in the locus coeruleus (LC), and then spreads to the entorhinal cortex, hippocampus, and finally, in advanced stages, throughout the frontal cortex. 15 , 282 , 283 , 284 Similar to the spread of Aβ pathology, TAU pathology also propagates to anatomically connected regions of the brain in a prion‐like manner: TAU aggregates either synthetically prepared or derived from AD‐affected brains trigger pathological TAU aggregation in mouse models of AD. 285 , 286 , 287 , 288 , 289 , 290 , 291 The self‐assembly of TAU into larger oligomers and filaments significantly depends on the presence of two hexapeptide motifs, PHF6 (306VQIVYK311) and PHF6*(275VQIINK280), within the MT‐binding domain. Deletion of the PHF6 sequence, which enables the assembly of TAU into β‐sheet conformations similar to that of TAU fibrils observed in vivo, for example is sufficient to prevent TAU assembly. 292 TAU aggregation can be further enhanced by the interaction with RNA and RNPs, as interaction of TAU with the RNP TIA1 for example increases stress granule formation and toxic TAU aggregation. 166
Normally, TAU is efficiently sorted into the axon. In the presence of pathological Aβ species, however, TAU undergoes abnormal phosphorylation, detaches from MTs, and accumulates in the cell body and dendrites (Figure 2B). 25 , 252 , 293 , 294 , 295 , 296 , 297 Specifically dendritic TAU has been implemented as an early driver of Aβ‐induced neurotoxicity: TAU locally destabilizes dendritic MTs by recruiting Tubulin‐Tyrosin‐Ligase‐Like‐6 (TTLL6s). TTLL6 polyglutamylates dendritic MTs, which triggers spastin‐mediated MT severing and causes synaptic and specifically dendritic spine dysfunction. 101 , 103 , 298 , 299 Furthermore, dendritic TAU recruits Fyn to the postsynaptic NMDA receptor complex, which causes the phosphorylation of NMDA receptor subunits and ultimately results in excitotoxicity 164 (reviewed in 156 ). PTMs of TAU, such as phosphorylation and acetylation increase the dendritic localization of TAU and influence its aggregation propensity, contributing to local MT destabilization and synapse dysfunction. 134 , 156 , 300
As a MT‐associated protein, TAU plays a significant role in regulating the transport of proteins and organelles along the axon by influencing cytoskeletal motor proteins such as dynein and kinesin. 301 PTMs of TAU result in a reduced MT‐binding affinity, detachment from axonal MTs, and lead to altered MT dynamics. Pathological TAU directly impedes the binding of motor proteins to axonal MTs, leading to deficient axonal transport and the accumulation of organelles and axonal proteins in the cell body (Figure 2B). 211 Breakdown of axonal transport, together with local disruption of MT dynamics, directly contributes to the synaptic impairments observed in AD. The therapeutic benefit of MT‐stabilizing drugs as demonstrated in preclinical AD models further supports the importance of TAU‐mediated MT dynamics in AD. 302 , 303 , 304 , 305 However, it may be more appropriate to correct the imbalanced ratio between labile and stable microtubules rather than solely concentrating on stabilizing MTs (reviewed, e.g., in 102 ). In sum, TAU aggregation and loss of physiological TAU function contribute significantly to the synaptic dysfunction and neuronal cell death observed in AD through altered cytoskeleton dynamics and axonal transport deficits (Figure 2B). 25 , 211 , 240
7.3. The contribution of TAU isoforms to AD pathology
While the role of TAU as a key mediator of Aβ‐induced neurotoxicity is well established, the contribution of individual isoforms remains largely unexplored. 143 Studies performed in disease context often focus on one specific TAU isoform, mainly 2N4R (= “full‐length TAU”), and make use of mutated TAU variants (associated with FTD) to induce TAU pathology. In WT mice, the 2N4R‐TAU interactome has been specifically linked to neurological diseases. 143 14 . In the human brain, however, 1N‐TAU isoforms are the predominantly expressed isoforms, that make up to 50% of the whole TAU protein expression. 37 While TAU depletion protects neurons and mice from AD pathology, re‐expression of 1N4R‐TAU, but not the other human TAU isoforms, restored the vulnerability of MAPT KO hiPSC‐derived neurons to AβO‐induced neuronal dysfunction. 100 In primary neurons, KD of 2N4R‐TAU by RNAi or a 2N‐TAU specific antibody specifically prevented AβO‐induced synaptotoxicity, suggesting that specific TAU isoforms might be sufficient to drive neurodegeneration in these model systems. 138 Overexpression of the human 0N4R and 0N3R TAU isoforms in Drosophila melanogaster resulted in a reduced lifespan upon 0N3R expression, but also lead to greater neurodegeneration upon 0N4R expression, suggesting that upon overexpression, and due to different interactions, TAU isoforms exhibit a differential toxicity. 238 Furthermore, a comprehensive study performed in flies characterized the effect of individual human TAU isoform overexpression and demonstrated that overall lifespan is affected equally by all isoforms. Interestingly, specifically 4R TAU isoforms (0N, 1N, and 2N) lead to significant learning and memory deficits, that were not observed in 3R expressing animals. 306 TAU has been shown to interact with RNPs and ribosomal subunits (reviewed in 142 ), and TAU oligomers formed by 1N4R induce a translational stress response and inhibit translation, resulting in stress granule formation and changes of dendrite length and morphology that further drive neuronal dysfunction in primary murine neurons. 163 , 170 , 307 , 308 Besides the likely contribution of the N‐terminal inserts to disease pathology, the presence of an additional repeat domain has been shown to enhance the aggregation propensity of TAU isoforms in vitro. The additional repeat domain of 4R TAU isoforms results in faster assembly into oligomers, which is further enhanced by the presence of additional N‐terminal inserts, present in 1N‐ and 2N‐TAU isoforms that reduce the critical concentration to form fibrils in vitro. 112 , 223 Furthermore, due to the additional repeat domain present in 4R‐TAU isoforms, these isoforms can be subjected to a higher number of PTMs, such as phosphorylation. Especially phosphorylation of the KxGS motifs, that are located within the MT‐binding domain, could further contribute to local disruption of MT dynamics and drive synapse loss in AD. 101 Furthermore, TAU spreading or seeding could be isoform‐specific: A recent study demonstrated that exposure of human iPSC‐derived astrocytes to extracellular vesicles isolated from 1N3R and 1N4R overexpressing primary rat neurons lead to similar uptake of 3R and 4R TAU; however, 3R‐containing vesicles were more toxic to the cells compared to 4R‐containing extracellular vesicles (EVs). 309 As TAU spreading has been shown to be mediated (among other pathways) via EVs (reviewed, e.g., in 310 ), these initial results support our hypothesis that rather individual TAU isoforms mediate TAU toxicity in disease context.
In conclusion, studies investigating the contribution of individual TAU isoforms to AD pathology remain sparse. Available results suggest a significant contribution of the 1N‐ and 2N4R‐TAU isoforms to AD pathology; however, in vivo validations of these experiments are needed to gain a better understanding of these processes. Clarifying the specific contributions of individual TAU isoforms will be critical for unraveling the intricate mechanisms underlying TAU‐related neurodegenerative disorders.
8. TAU‐TARGETING THERAPIES FOR AD AND OTHER TAUOPATHIES
Among the seven currently U.S. Food and Drug Administration (FDA) ‐approved drugs for the treatment of AD, five (donepezil, rivastigmine, galantamine, memantine, and the combination of memantine with donepezil) primarily target disease symptoms, such as behavioral psychological, and cognitive impairments (e.g., agitation, aggression, memory loss), without substantially improving cognition or halting disease progression. 239 Two antibodies designed to target Aβ, aducanumab (Biogen) and lecanemab (Eisai and Biogen), have gained approval as disease‐modifying treatments (DMTs). Both drugs received FDA approval via the accelerated approval pathway, which relies on a “surrogate endpoint” anticipated to correlate with a positive impact on cognitive function. For full FDA approval, both medications must demonstrate their clinical efficacy in subsequent trials.
With respect to TAU, several agents are currently investigated in preclinical and clinical settings, which use diverse strategies to diminish or avert pathological TAU build‐up in the context of AD and other tauopathies (for an overview of current clinical trials for AD, see 311 ). For example, reducing overall MAPT expression by an antisense oligonucleotide (ASO) significantly reduced TAU pathology and improved cognition in AD/tauopathy mouse models and has proven safety in a phase Ib clinical study. 277 , 312 , 313 The ASO, termed MAPTRx (Ionis Pharmaceuticals), binds to intron 9 of the MAPT pre‐mRNA, thereby targeting it for degradation and achieving up to 50% reduction of TAU levels in the cerebrospinal fluid (CSF) of human subjects. 312 A successful and significant reduction of TAU expression and neuronal degeneration has been achieved also by a single systemic injection of an AAV‐based engineered zinc finger protein transcription factor in APP/PS1 mice. 314 While some antibody‐based therapies for the clearance of extracellular Aβ have shown clinical benefit and have been approved for use in AD patients, no effective therapy targeting TAU has yet shown efficacy. Several antibodies are currently investigated in clinical studies to target toxic TAU species, such as highly phosphorylated and aggregation‐prone TAU or TAU assemblies. Unlike Aβ, the majority of (toxic) TAU species may remain intracellularly, and it needs to be demonstrated whether antibodies can successfully prevent pathology and improve cognition by targeting mostly extracellular TAU species (for review, see e.g., 315 , 316 ). Besides enhancing the clearance of pathological TAU, several compounds have been designed to inhibit TAU aggregation. Leuco‐Methylthioninium Bis(Hydromethanesulphonate) (LMTM/ TauRx Pharmaceuticals), a derivative of methylene blue, has demonstrated a significant reduction of TAU pathology. In TAU transgenic mice, LMTM improved cognitive impairments, 317 and a recent phase III study on early AD patients demonstrated an improvement of cognitive deficits and a reduction of brain atrophy. 316 , 318 , 319 ACI‐3024 (AC Immune) is another small compound TAU aggregation inhibitor. Even though preclinical data have not been published yet, data presented at conferences show that the compound disrupts TAU aggregates in vitro, in primary neuronal cultures and in a tauopathy mouse model 320 (as discussed in 316 ). Further, it prevents microglial activation and neuronal cell death induced by toxic TAU species. A phase I study was started in 2019 to evaluate the safety and tolerability of ACI‐3024 in healthy volunteers, however results have not yet been reported (reviewed in 316 ). Other therapeutic strategies aim to target specific PTMs of TAU, for example, (1) by inhibiting GSK3β and preventing (hyper)phosphorylation of TAU 66 , 321 , 322 ; (2) by targeting acetylation of TAU, which has been demonstrated to increase the aggregation propensity of TAU 134 , 156 , 300 ; or (3) by inhibiting O‐GlcNAcase (OGA) activity, to increase O‐GlcNAcetylation of TAU, which has been reported to prevent aggregation of TAU, and stabilize it in a soluble form. 323 , 324 While promising results have been demonstrated in preclinical models for GSK3β inhibitors Lithium, valproic acid, and Tideglusib, 66 , 321 , 322 , 325 none of the compounds demonstrated beneficial effects on cognition in clinical trials. 326 , 327 , 328 Salsalate, which has been found to inhibit TAU acetylation in animal models for FTD and TBI, did not show any clinical benefit in PSP patients. 329 , 330 , 331 While inhibiting phosphorylation and acetylation of TAU did not demonstrate beneficial outcomes in clinical studies so far, three OGA activity inhibitors, MK‐8719 (Alectos Therapeutics/ Merck), ASN90 (Asceneuron SA), and LY3884963 (Eli Lilly & Co) received orphan drug designation by the FDA for the treatment of PSP (MK‐8719, ASN90), and FTD (LY3884963). 332 , 333 , 334
Despite the variety of TAU‐targeting therapeutic strategies, to our knowledge, no DMT for AD or other TAU‐related neurodegenerative disease that specifically targets a particular TAU isoform or alters MAPT splicing is currently being evaluated in preclinical or clinical trials for AD. While an imbalance in TAU 3R‐to‐4R isoform expression has been directly linked to disease pathogenesis in several tauopathies, 9 , 10 , 207 , 212 data on the contribution of individual TAU isoforms on AD are scarce. Nevertheless, some recent results imply benefits in disease models when individual isoforms are suppressed, or demonstrate isoform‐specific mediation of neuronal dysfunction. 100 , 138 (see the The contribution of TAU isoforms to AD pathology section) A strategy to enhance exon 10 incorporation through a lentivirus‐mediated trans‐splicing event has shown beneficial results in the humanized TAU mouse model of Andorfer and colleagues, restoring equal ratios of 3R and 4R TAU expression and reducing TAU pathology and cognitive impairments. 191 , 335 , 336 , 337 Targeting specific TAU isoforms in AD may be advantageous over total protein reduction strategies, as total TAU reduction can lead to mild age‐related phenotypes in animal models and is associated with a severe developmental disorder in humans (see the The effect of TAU depletion and (re‐)expression of (individual) TAU isoforms section).
9. CONCLUSION AND PERSPECTIVE
The species‐dependent and tissue‐specific complexity of MAPT splicing and the resulting diversity of TAU isoforms may have evolutionary advantages: Based on studies on TAU isoforms performed in murine and human model systems, we can assume that the isoforms adopt distinct functions in dependence of, for example, their expression timing during development, brain‐region specific expression, PTM‐sites, interactors, and subcellular localization.
While MAPT splicing regulation may provide evolutionary advantages in a physiological context as outlined above, their regulation in humans apparently has to be maintained in a tight fashion: Dysregulation of MAPT splicing, altered subcellular localizations, and differences in PTMs influencing the aggregation propensity of the protein significantly contribute to disease pathogenesis and neuronal dysfunction, all proven by clinical disease manifestation or experimental evidence. The failure of a variety of Aβ‐ and TAU‐targeting therapies in AD, the most prevalent tauopathy, highlights the urgent need for better targeted therapies. The following points should be considered: (1) Therapeutic strategies targeting TAU isoforms, including splicing modifiers and interventions that restore the correct isoform ratio, hold promise as effective therapies for TAU‐related diseases, in particular for diseases where TAU isoform balance per se is affected; (2) The development of TAU‐based (independent of whether strategies are isoform‐based or not) therapeutic approaches must take into consideration that MAPT splicing patterns show developmental stage‐, region‐, and cell type‐ specificity in the brain, suggesting selective susceptibility to changes in TAU isoform expression, and that the role of TAU in neuronal development needs to be carefully assessed, as 0N3R‐TAU may be essential for proper human brain development, which would imply also an important function for adult neurogenesis; and finally, (3) Future disease models of Alzheimer's or other tauopathies should be developed on the basis of systems that enable the study of the different TAU isoforms, for example, mice or human neurons expressing all human TAU isoforms. Better understanding of TAU biology and the involvement of the TAU isoforms in physiology and disease mechanisms will ultimately help to develop specific therapies, which must target the pathological processes underlying tauopathies, all while preserving the essential functions of TAU in normal cellular physiology.
CONFLICT OF INTEREST STATEMENT
The authors declare no competing interest. Author disclosures are available in the supporting information.
Supporting information
Supporting information
ACKNOWLEDGMENTS
The authors thank Michael Bell‐Simons for critical manuscript review and David Benitez Alvarez for help with the revision. The authors acknowledge DFG (German Research Foundation, 491454339) for support of the Article Processing Charge. Supported by the Alzheimer‐Forschungs‐Initiative (grant number 22039) and a publication grant by the Alzheimer‐Forschungs‐Initiative to Sarah Buchholz.
Open access funding enabled and organized by Projekt DEAL.
Buchholz S, Zempel H. The six brain‐specific TAU isoforms and their role in Alzheimer's disease and related neurodegenerative dementia syndromes. Alzheimer's Dement. 2024;20:3606–3628. 10.1002/alz.13784
REFERENCES
- 1. Barbier P, Zejneli O, Martinho M, et al. Role of tau as a microtubule‐associated protein: structural and functional aspects. Front Aging Neurosci. 2019;10:204. doi: 10.3389/FNAGI.2019.00204/BIBTEX [DOI] [PMC free article] [PubMed] [Google Scholar]
- 2. Goedert M, Spillantini MG, Jakes R, Rutherford D, Crowther RA. Multiple isoforms of human microtubule‐associated protein tau: sequences and localization in neurofibrillary tangles of Alzheimer's disease. Neuron. 1989. doi: 10.1016/0896-6273(89)90210-9 [DOI] [PubMed] [Google Scholar]
- 3. Andreadis A, Brown WM, Kosik KS. Structure and novel exons of the human tau gene. Biochemistry. 1992;31:10626‐10633. doi: 10.1021/bi00158a027 [DOI] [PubMed] [Google Scholar]
- 4. Couchie D, Mavilia C, Georgieff IS, Liem RK, Shelanski ML, Nunez J. Primary structure of high molecular weight tau present in the peripheral nervous system. Proc Natl Acad Sci. 1992;89:4378‐4381. doi: 10.1073/pnas.89.10.4378 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 5. Goedert M, Spillantini MG, Crowther RA. Cloning of a big tau microtubule‐associated protein characteristic of the peripheral nervous system. Proc Natl Acad Sci. 1992;89:1983‐1987. doi: 10.1073/pnas.89.5.1983 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 6. Janke C, Beck M, Stahl T, Holzer M, Brauer K, Bigl V, et al. Phylogenetic diversity of the expression of the microtubule‐associated protein tau: implications for neurodegenerative disorders. Mol Brain Res. 1999;68:119‐128. doi: 10.1016/S0169-328X(99)00079-0 [DOI] [PubMed] [Google Scholar]
- 7. Caillet‐Boudin ML, Buée L, Sergeant N, Lefebvre B. Regulation of human MAPT gene expression. Mol Neurodegener. 2015;10. doi: 10.1186/S13024-015-0025-8 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 8. Goedert M, Spillantini MG, Potier MC, Ulrich J, Crowther RA. 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. 1989;8:393‐399. doi: 10.1002/j.1460-2075.1989.tb03390.x [DOI] [PMC free article] [PubMed] [Google Scholar]
- 9. Ling H. Untangling the tauopathies: current concepts of tau pathology and neurodegeneration. Parkinsonism Relat Disord. 2018;46:S34‐38. doi: 10.1016/j.parkreldis.2017.07.031 [DOI] [PubMed] [Google Scholar]
- 10. Zhang Y, Wu KM, Yang L, Dong Q, Yu JT. Tauopathies: new perspectives and challenges. Mol Neurodegener. 2022;17:1‐29. doi: 10.1186/S13024-022-00533-Z [DOI] [PMC free article] [PubMed] [Google Scholar]
- 11. Scheltens P, De Strooper B, Kivipelto M, Holstege H, Chételat G, Teunissen CE, et al. Alzheimer's disease. Lancet. 2021;397:1577‐1590. doi: 10.1016/S0140-6736(20)32205-4 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 12. Long JM, Holtzman DM. Alzheimer disease: an update on pathobiology and treatment strategies. Cell. 2019;179:312‐339. doi: 10.1016/j.cell.2019.09.001 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 13. Knopman DS, Amieva H, Petersen RC, Chételat G, Holtzman DM, Hyman BT, et al. Alzheimer disease. Nat Rev Dis Primers. 2021;7:33. doi: 10.1038/s41572-021-00269-y [DOI] [PMC free article] [PubMed] [Google Scholar]
- 14. Götz J, Chen F, Van Dorpe J, Nitsch RM. Formation of neurofibrillary tangles in P301L tau transgenic mice induced by Aβ42 fibrils. Science (1979). 2001;293:1491‐1495. doi: 10.1126/SCIENCE.1062097/ASSET/2313AA72-528B-4FB8-9641-CD8E0B30E73F/ASSETS/GRAPHIC/SE3319696003.JPEG [DOI] [PubMed] [Google Scholar]
- 15. Braak H, Braak E. Neuropathological stageing of Alzheimer‐related changes. Acta Neuropathol. 1991;82:239‐259. doi: 10.1007/BF00308809 [DOI] [PubMed] [Google Scholar]
- 16. del Alonso AC, Mederlyova A, Novak M, Grundke‐Iqbal I, Iqbal K. Promotion of hyperphosphorylation by frontotemporal dementia tau mutations. J Biol Chem. 2004;279:34873‐34881. doi: 10.1074/jbc.M405131200 [DOI] [PubMed] [Google Scholar]
- 17. Zempel H. Genetic and sporadic forms of tauopathies–TAU as a disease driver for the majority of patients but the minority of tauopathies. Cytoskeleton. 2023. doi: 10.1002/CM.21793 [DOI] [PubMed] [Google Scholar]
- 18. Spillantini MG, Murrell JR, Goedert M, Farlow MR, Klug A, Ghetti B. Mutation in the tau gene in familial multiple system tauopathy with presenile dementia. Proc Natl Acad Sci USA. 1998;95:7737‐7741. doi: 10.1073/PNAS.95.13.7737 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 19. Stanford PM, Halliday GM, Brooks WS, et al. Progressive supranuclear palsy pathology caused by a novel silent mutation in exon 10 of the tau gene: expansion of the disease phenotype caused by tau gene mutations. Brain. 2000;123(5):880‐893. doi: 10.1093/BRAIN/123.5.880 [DOI] [PubMed] [Google Scholar]
- 20. Stanford PM, Shepherd CE, Halliday GM, et al. Mutations in the tau gene that cause an increase in three repeat tau and frontotemporal dementia. Brain. 2003;126:814‐826. doi: 10.1093/BRAIN/AWG090 [DOI] [PubMed] [Google Scholar]
- 21. Gauthier‐Kemper A, Weissmann C, Golovyashkina N, et al. The frontotemporal dementia mutation R406W blocks tau's interaction with the membrane in an annexin A2‐dependent manner. J Cell Biol. 2011. doi: 10.1083/jcb.201007161 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 22. Bang J, Spina S, Miller BL. Frontotemporal dementia. Lancet North Am Ed. 2015;386:1672‐1682. doi: 10.1016/S0140-6736(15)00461-4 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 23. Greaves CV, Rohrer JD. An update on genetic frontotemporal dementia. J Neurol. 2019;266:2075‐2086. doi: 10.1007/S00415-019-09363-4/FIGURES/3 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 24. Buée L, Delacourte A. Comparative biochemistry of tau in progressive supranuclear palsy, corticobasal degeneration, FTDP‐17 and Pick's disease. Brain Pathol. 1999;9:681‐693. doi: 10.1111/J.1750-3639.1999.TB00550.X [DOI] [PMC free article] [PubMed] [Google Scholar]
- 25. Arendt T, Stieler JT, Holzer M. Tau and tauopathies. Brain Res Bull. 2016;126:238‐292. doi: 10.1016/j.brainresbull.2016.08.018 [DOI] [PubMed] [Google Scholar]
- 26. Lopresti P, Szuchet S, Papasozomenos SC, Zinkowski RP, Binder LI. Functional implications for the microtubule‐associated protein tau: localization in oligodendrocytes. Proc Natl Acad Sci. 1995;92:10369‐10373. doi: 10.1073/PNAS.92.22.10369 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 27. Kubo A, Misonou H, Matsuyama M, et al. Distribution of endogenous normal tau in the mouse brain. J Comp Neurol. 2019;527:985‐998. doi: 10.1002/CNE.24577 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 28. Ezerskiy LA, Schoch KM, Sato C, et al. Astrocytic 4R tau expression drives astrocyte reactivity and dysfunction. JCI Insight. 2022;7. doi: 10.1172/JCI.INSIGHT.152012 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 29. McKenzie AT, Wang M, Hauberg ME, et al. Brain cell type specific gene expression and co‐expression network architectures. Sci Rep. 2018;8:1‐19. doi: 10.1038/s41598-018-27293-5 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 30. Darmanis S, Sloan SA, Zhang Y, et al. A survey of human brain transcriptome diversity at the single cell level. Proc Natl Acad Sci USA. 2015;112:7285‐7290. doi: 10.1073/PNAS.1507125112/SUPPL_FILE/PNAS.1507125112.SAPP.PDF [DOI] [PMC free article] [PubMed] [Google Scholar]
- 31. Seiberlich V, Bauer NG, Schwarz L, Ffrench‐Constant C, Goldbaum O, Richter‐Landsberg C. Downregulation of the microtubule associated protein Tau impairs process outgrowth and myelin basic protein mRNA transport in oligodendrocytes. Glia. 2015;63:1621‐1635. doi: 10.1002/GLIA.22832 [DOI] [PubMed] [Google Scholar]
- 32. Zhang Y, Chen K, Sloan SA, et al. An RNA‐sequencing transcriptome and splicing database of glia, neurons, and vascular cells of the cerebral cortex. J Neurosci. 2014;34:11929‐11947. doi: 10.1523/JNEUROSCI.1860-14.2014 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 33. Corsi A, Bombieri C, Valenti MT, Romanelli MG. Tau isoforms: gaining insight into MAPT alternative splicing. Int J Mol Sci. 2022;23:15383. doi: 10.3390/IJMS232315383 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 34. Gu Y, Oyama F, Ihara Y. τ is widely expressed in rat tissues. J Neurochem. 1996;67:1235‐1244. doi: 10.1046/J.1471-4159.1996.67031235.X [DOI] [PubMed] [Google Scholar]
- 35. Arikan MC, Memmott J, Broderick JA, et al. Modulation of the membrane‐binding projection domain of tau protein: splicing regulation of exon 3. Mol Brain Res. 2002;101:109‐121. doi: 10.1016/S0169-328X(02)00178-X [DOI] [PubMed] [Google Scholar]
- 36. Hefti MM, Farrell K, Kim S, et al. High‐resolution temporal and regional mapping of MAPT expression and splicing in human brain development. PLoS One. 2018;13:e0195771. doi: 10.1371/journal.pone.0195771 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 37. Trabzuni D, Wray S, Vandrovcova J, et al. MAPT expression and splicing is differentially regulated by brain region: relation to genotype and implication for tauopathies. Hum Mol Genet. 2012;21:4094‐4103. doi: 10.1093/HMG/DDS238 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 38. McMillan P, Korvatska E, Poorkaj P, et al. Tau isoform regulation is region‐ and cell‐specific in mouse brain. J Comp Neurol. 2008;511:788‐803. doi: 10.1002/cne.21867 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 39. Boutajangout A, Boom A, Leroy K, Brion JP. Expression of tau mRNA and soluble tau isoforms in affected and non‐affected brain areas in Alzheimer's disease. FEBS Lett. 2004;576:183‐189. doi: 10.1016/j.febslet.2004.09.011 [DOI] [PubMed] [Google Scholar]
- 40. Majounie E, Cross W, Newsway V, et al. Variation in tau isoform expression in different brain regions and disease states. Neurobiol Aging. 2013;34:1922.e7‐1922.e12. doi: 10.1016/j.neurobiolaging.2013.01.017 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 41. Fischer I, Baas PW. Resurrecting the mysteries of big tau. Trends Neurosci. 2020;43:493‐504. doi: 10.1016/j.tins.2020.04.007 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 42. Fischer I. Big tau: what we know, and we need to know. ENeuro. 2023;10. doi: 10.1523/ENEURO.0052-23.2023 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 43. Dugger BN, Whiteside CM, Maarouf CL, et al. The presence of select tau species in human peripheral tissues and their relation to Alzheimer's disease. J Alzheimers Dis. 2016;51:345‐356. doi: 10.3233/JAD-150859 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 44. Dugger BN, Hoffman BR, Scroggins A, et al. Tau immunoreactivity in peripheral tissues of human aging and select tauopathies. Neurosci Lett. 2019;696:132‐139. doi: 10.1016/J.NEULET.2018.12.031 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 45. Lionnet A, Wade MA, Corbillé AG, et al. Characterisation of tau in the human and rodent enteric nervous system under physiological conditions and in tauopathy. Acta Neuropathol Commun. 2018;6:1‐17. doi: 10.1186/S40478-018-0568-3/FIGURES/8 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 46. Maj M, Gartner W, Ilhan A, Neziri D, Attems J, Wagner L. Expression of TAU in insulin‐secreting cells and its interaction with the calcium‐binding protein secretagogin. J Endocrinol. 2010;205:25‐36. doi: 10.1677/JOE-09-0341 [DOI] [PubMed] [Google Scholar]
- 47. Zhou R, Hu W, Dai CL, et al. Expression of microtubule associated protein tau in mouse pancreatic islets is restricted to autonomic nerve fibers. J Alzheimers Dis. 2020;75:1361‐1376. doi: 10.3233/JAD-200101 [DOI] [PubMed] [Google Scholar]
- 48. Niblock M, Gallo JM. Tau alternative splicing in familial and sporadic tauopathies. Biochem Soc Trans. 2012;40:677‐680. doi: 10.1042/BST20120091 [DOI] [PubMed] [Google Scholar]
- 49. Park SA, Ahn S Il, Gallo JM. Tau mis‐splicing in the pathogenesis of neurodegenerative disorders. BMB Rep. 2016;49:405. doi: 10.5483/BMBREP.2016.49.8.084 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 50. Andreadis A. Tau gene alternative splicing: expression patterns, regulation and modulation of function in normal brain and neurodegenerative diseases. Biochim Biophys Acta Mol Basis Dis. 2005;1739:91‐103. doi: 10.1016/J.BBADIS.2004.08.010 [DOI] [PubMed] [Google Scholar]
- 51. Gao QS, Memmott J, Lafyatis R, Stamm S, Screaton G, Andreadis A. Complex regulation of tau exon 10, whose missplicing causes frontotemporal dementia. J Neurochem. 2000;74:490‐500. doi: 10.1046/J.1471-4159.2000.740490.X [DOI] [PubMed] [Google Scholar]
- 52. Ghetti B, Oblak AL, Boeve BF, Johnson KA, Dickerson BC, Goedert M. Invited review: frontotemporal dementia caused by microtubule‐associated protein tau gene (MAPT) mutations: a chameleon for neuropathology and neuroimaging. Neuropathol Appl Neurobiol. 2015;41:24‐46. doi: 10.1111/NAN.12213 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 53. Kowalska A, Hasegawa M, Miyamoto K, et al. A novel mutation at position +11 in the intron following exon 10 of the tau gene in FTDP‐17. J Appl Genet. 2002;43:535‐543. [PubMed] [Google Scholar]
- 54. Malkani R, D'Souza I, Gwinn‐Hardy K, Schellenberg GD, Hardy J, Momeni P. A MAPT mutation in a regulatory element upstream of exon 10 causes frontotemporal dementia. Neurobiol Dis. 2006;22:401‐403. doi: 10.1016/J.NBD.2005.12.001 [DOI] [PubMed] [Google Scholar]
- 55. Anfossi M, Vuono R, Maletta R, et al. Compound heterozygosity of 2 novel MAPT mutations in frontotemporal dementia. Neurobiol Aging. 2011;32:757.e1‐757.e11. doi: 10.1016/J.NEUROBIOLAGING.2010.12.013 [DOI] [PubMed] [Google Scholar]
- 56. McCarthy A, Lonergan R, Olszewska DA, et al. Closing the tau loop: the missing tau mutation. Brain. 2015;138:3100‐3109. doi: 10.1093/BRAIN/AWV234 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 57. Qian W, Liang H, Shi J, et al. Regulation of the alternative splicing of tau exon 10 by SC35 and Dyrk1A. Nucleic Acids Res. 2011;39:6161‐6171. doi: 10.1093/NAR/GKR195 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 58. D'Souza I, Poorkaj P, Hong M, et al. Missense and silent tau gene mutations cause frontotemporal dementia with parkinsonism‐chromosome 17 type, by affecting multiple alternative RNA splicing regulatory elements. Proc Natl Acad Sci USA. 1999;96:5598‐5603. doi: 10.1073/PNAS.96.10.5598/ASSET/B6525F05-ECE5-48F6-85E5-C911F3D2205C/ASSETS/GRAPHIC/PQ1090982005.JPEG [DOI] [PMC free article] [PubMed] [Google Scholar]
- 59. Clark LN, Poorkaj P, Wszolek Z, et al. Pathogenic implications of mutations in the tau gene in pallido‐ponto‐nigral degeneration and related neurodegenerative disorders linked to chromosome 17. Proc Natl Acad Sci USA. 1998;95:13103‐13107. doi: 10.1073/PNAS.95.22.13103 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 60. Rizzu P, Van Swieten JC, Joosse M, et al. High prevalence of mutations in the microtubule‐associated protein tau in a population study of frontotemporal dementia in the Netherlands. Am J Hum Genet. 1999;64:414‐421. doi: 10.1086/302256 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 61. Donahue CP, Muratore C, Wu JY, Kosik KS, Wolfe MS. Stabilization of the tau exon 10 stem loop alters pre‐mRNA splicing. J Biol Chem. 2006;281:23302‐23306. doi: 10.1074/JBC.C600143200 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 62. Lisowiec J, Magner D, Kierzek E, Lenartowicz E, Kierzek R. Structural determinants for alternative splicing regulation of the MAPT pre‐mRNA. RNA Biol. 2015;12:330. doi: 10.1080/15476286.2015.1017214 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 63. Yin X, Jin N, Gu J, et al. Dual‐specificity tyrosine phosphorylation‐regulated kinase 1A (Dyrk1A) modulates serine/arginine‐rich protein 55 (SRp55)‐promoted tau exon 10 inclusion. J Biol Chem. 2012;287:30497‐30506. doi: 10.1074/JBC.M112.355412 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 64. Chen C, Jin N, Qian W, et al. Cyclic AMP‐dependent protein kinase enhances SC35‐promoted Tau Exon 10 inclusion. Mol Neurobiol. 2014;49:615‐624. doi: 10.1007/S12035-013-8542-3/FIGURES/6 [DOI] [PubMed] [Google Scholar]
- 65. Hartmann AM, Rujescu D, Giannakouros T, et al. Regulation of alternative splicing of human tau exon 10 by phosphorylation of splicing factors. Mol Cell Neurosci. 2001;18:80‐90. doi: 10.1006/MCNE.2001.1000 [DOI] [PubMed] [Google Scholar]
- 66. Noble W, Planel E, Zehr C, et al. Inhibition of glycogen synthase kinase‐3 by lithium correlates with reduced tauopathy and degeneration in vivo. Proc Natl Acad Sci USA. 2005;102. doi: 10.1073/pnas.0500466102 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 67. Dredge BK, Stefani G, Engelhard CC, Darnell RB. Nova autoregulation reveals dual functions in neuronal splicing. EMBO J. 2005;24:1608‐1620. doi: 10.1038/SJ.EMBOJ.7600630/ASSET/DDFF938B-CEAC-47D6-B795-00D318782D25/ASSETS/GRAPHIC/EMBJ7600630-FIG-0008-M.JPG [DOI] [PMC free article] [PubMed] [Google Scholar]
- 68. López‐Martínez A, Soblechero‐Martín P, De‐La‐puente‐ovejero L, Nogales‐Gadea G, Arechavala‐Gomeza V. An overview of alternative splicing defects implicated in myotonic dystrophy type I. Genes. 2020;11:1109. doi: 10.3390/GENES11091109 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 69. Degener MJF, Van Cruchten RTP, Otero BA, Wang ET, Wansink DG, Hoen PAC. A comprehensive atlas of fetal splicing patterns in the brain of adult myotonic dystrophy type 1 patients. NAR Genom Bioinform. 2022;4. doi: 10.1093/NARGAB/LQAC016 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 70. Bazrgar M, Khodabakhsh P, Mohagheghi F, Prudencio M, Ahmadiani A. Brain microRNAs dysregulation: implication for missplicing and abnormal post‐translational modifications of tau protein in Alzheimer's disease and related tauopathies. Pharmacol Res. 2020;155. doi: 10.1016/J.PHRS.2020.104729 [DOI] [PubMed] [Google Scholar]
- 71. Praticò D. The functional role of microRNAs in the pathogenesis of tauopathy. Cells. 2020;9. doi: 10.3390/CELLS9102262 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 72. Wang L, Shui X, Diao Y, Chen D, Zhou Y, Lee TH. Potential implications of miRNAs in the pathogenesis, diagnosis, and therapeutics of Alzheimer's disease. Int J Mol Sci. 2023;24:16259. doi: 10.3390/IJMS242216259 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 73. Ruiz‐Gabarre D, Carnero‐Espejo A, Ávila J, García‐Escudero V. What's in a gene? The outstanding diversity of MAPT. 2022:840. doi: 10.3390/CELLS11050840 [DOI] [PMC free article] [PubMed]
- 74. Liu F, Gong CX. Tau exon 10 alternative splicing and tauopathies. Mol Neurodegener. 2008;3:8. doi: 10.1186/1750-1326-3-8 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 75. Andreadis A. Tau splicing and the intricacies of dementia. J Cell Physiol. 2012;227:1220‐1225. doi: 10.1002/jcp.22842 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 76. Caffrey TM, Wade‐Martins R. Functional MAPT haplotypes: bridging the gap between genotype and neuropathology. Neurobiol Dis. 2007;27:1‐10. doi: 10.1016/J.NBD.2007.04.006 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 77. Pittman AM, Myers AJ, Abou‐Sleiman P, et al. Linkage disequilibrium fine mapping and haplotype association analysis of the tau gene in progressive supranuclear palsy and corticobasal degeneration. J Med Genet. 2005;42:837‐846. doi: 10.1136/JMG.2005.031377 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 78. Caffrey TM, Wade‐Martins R. The role of MAPT sequence variation in mechanisms of disease susceptibility. Biochem Soc Trans. 2012;40:687‐692. doi: 10.1042/BST20120063 [DOI] [PubMed] [Google Scholar]
- 79. Pastor P, Ezquerra M, Muñoz E, et al. Significant association between the tau gene A0/A0 genotype and Parkinson's disease. Ann Neurol. 2000;47:242‐245. doi: [DOI] [PubMed] [Google Scholar]
- 80. Di Maria E, Tabaton M, Vigo T, et al. Corticobasal degeneration shares a common genetic background with progressive supranuclear palsy. Ann Neurol. 2000;47:374‐377. doi: [DOI] [PubMed] [Google Scholar]
- 81. Caffrey TM, Joachim C, Paracchini S, Esiri MM, Wade‐Martins R. Haplotype‐specific expression of exon 10 at the human MAPT locus. Hum Mol Genet. 2006;15:3529‐3537. doi: 10.1093/hmg/ddl429 [DOI] [PubMed] [Google Scholar]
- 82. Stefansson H, Helgason A, Thorleifsson G, et al. A common inversion under selection in Europeans. Nat Genet. 2005;37:129‐137. doi: 10.1038/ng1508 [DOI] [PubMed] [Google Scholar]
- 83. Baker M, Litvan I, Houlden H, et al. Association of an extended haplotype in the tau gene with progressive supranuclear palsy. Hum Mol Genet. 1999;8:711‐715. doi: 10.1093/HMG/8.4.711 [DOI] [PubMed] [Google Scholar]
- 84. Allen M, Kachadoorian M, Quicksall Z, et al. Association of MAPT haplotypes with Alzheimer's disease risk and MAPT brain gene expression levels. Alzheimers Res Ther. 2014;6:1‐14. doi: 10.1186/ALZRT268/TABLES/4 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 85. Evans W, Fung HC, Steele J, et al. The tau H2 haplotype is almost exclusively Caucasian in origin. Neurosci Lett. 2004;369:183‐185. doi: 10.1016/J.NEULET.2004.05.119 [DOI] [PubMed] [Google Scholar]
- 86. Takuma H, Arawaka S, Mori H. Isoforms changes of tau protein during development in various species. Dev Brain Res. 2003;142:121‐127. doi: 10.1016/S0165-3806(03)00056-7 [DOI] [PubMed] [Google Scholar]
- 87. Bullmann T, Holzer M, Mori H, Arendt T. Pattern of tau isoforms expression during development in vivo. Int J Dev Neurosci. 2009;27:591‐597. doi: 10.1016/J.IJDEVNEU.2009.06.001 [DOI] [PubMed] [Google Scholar]
- 88. Kampers T, Pangalos M, Geerts H, Wiech H, Mandelkow E. Assembly of paired helical filaments from mouse tau: implications for the neurofibrillary pathology in transgenic mouse models for Alzheimer's disease. FEBS Lett. 1999;451:39‐44. doi: 10.1016/S0014-5793(99)00522-0 [DOI] [PubMed] [Google Scholar]
- 89. Götz J. Tau and transgenic animal models. Brain Res Rev. 2001;35:266‐286. doi: 10.1016/S0165-0173(01)00055-8 [DOI] [PubMed] [Google Scholar]
- 90. Uversky VN. Intrinsically disordered proteins and their (disordered) proteomes in neurodegenerative disorders. Front Aging Neurosci. 2015;7. doi: 10.3389/FNAGI.2015.00018/FULL [DOI] [PMC free article] [PubMed] [Google Scholar]
- 91. Jeganathan S, Von Bergen M, Mandelkow EM, Mandelkow E. The natively unfolded character of Tau and its aggregation to Alzheimer‐like paired helical filaments. Biochemistry. 2008. doi: 10.1021/bi800783d [DOI] [PubMed] [Google Scholar]
- 92. Chudobová J, Zempel H. Microtubule affinity regulating kinase (MARK/Par1) isoforms differentially regulate Alzheimer‐like TAU missorting and Aβ‐mediated synapse pathology. Neural Regen Res. 2023;18:335‐336. doi: 10.4103/1673-5374.346477 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 93. Andronesi OC, Von Bergen M, Biernat J, et al. Characterization of Alzheimer's‐like paired helical filaments from the core domain of tau protein using solid‐state NMR spectroscopy. J Am Chem Soc. 2008;130:5922‐5928. doi: 10.1021/JA7100517/SUPPL_FILE/JA7100517-FILE003.PDF [DOI] [PubMed] [Google Scholar]
- 94. Mukrasch MD, Bibow S, Korukottu J, et al. Structural polymorphism of 441‐residue tau at single residue resolution. PLoS Biol. 2009;7:e1000034. doi: 10.1371/JOURNAL.PBIO.1000034 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 95. Jeganathan S, Hascher A, Chinnathambi S, Biernat J, Mandelkow EM, Mandelkow E. Proline‐directed pseudo‐phosphorylation at AT8 and PHF1 epitopes induces a compaction of the paperclip folding of tau and generates a pathological (MC‐1) conformation. J Biol Chem. 2008;283:32066‐32076. doi: 10.1074/JBC.M805300200 [DOI] [PubMed] [Google Scholar]
- 96. Hirokawa N, Shiomura Y, Okabe S. Tau proteins: the molecular structure and mode of binding on microtubules. J Cell Biol. 1988;107:1449‐1459. doi: 10.1083/JCB.107.4.1449 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 97. Qiang L, Sun X, Austin TO, et al. Tau does not stabilize axonal microtubules but rather enables them to have long labile domains. Curr Biol. 2018;28:2181‐2189. doi: 10.1016/j.cub.2018.05.045.e4 [DOI] [PubMed] [Google Scholar]
- 98. Rapoport M, Dawson HN, Binder LI, Vitek MP, Ferreira A. Tau is essential to beta ‐amyloid‐induced neurotoxicity. Proc Natl Acad Sci USA. 2002;99:6364‐6369. doi: 10.1073/pnas.092136199 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 99. Dawson HN, Ferreira A, Eyster MV, Ghoshal N, Binder LI, Vitek MP. Inhibition of neuronal maturation in primary hippocampal neurons from tau deficient mice. J Cell Sci. 2001. [DOI] [PubMed] [Google Scholar]
- 100. Buchholz S, Bell‐Simons M, Aghyad M, et al. The TAU isoform 1N4R restores vulnerability of MAPT knockout human iPSC‐derived neurons to Amyloid beta‐induced neuronal dysfunction. 2022. doi: 10.21203/RS.3.RS-2277268/V1 [DOI]
- 101. Zempel H, Luedtke J, Kumar Y, et al. Amyloid‐β oligomers induce synaptic damage via Tau‐dependent microtubule severing by TTLL6 and spastin. EMBO J. 2013;32:2920‐2937. doi: 10.1038/EMBOJ.2013.207 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 102. Baas PW, Qiang L. Tau: it's not what you think. Trends Cell Biol. 2019;29:452‐461. doi: 10.1016/J.TCB.2019.02.007 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 103. Zempel H, Mandelkow EM. Tau missorting and spastin‐induced microtubule disruption in neurodegeneration: Alzheimer disease and hereditary spastic paraplegia. Mol Neurodegener. 2015;10. doi: 10.1186/s13024-015-0064-1 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 104. Chen J, Kanai Y, Cowan NJ, Hirokawa N. Projection domains of MAP2 and tau determine spacings between microtubules in dendrites and axons. Nature. 1992. doi: 10.1038/360674a0 [DOI] [PubMed] [Google Scholar]
- 105. Drechsel DN, Hyman AA, Cobb MH, Kirschner MW. Modulation of the dynamic instability of tubulin assembly by the microtubule‐associated protein tau. Mol Biol Cell. 1992. doi: 10.1091/mbc.3.10.1141 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 106. Witman GB, Cleveland DW, Weingarten MD, Kirschner MW. Tubulin requires tau for growth into microtubule initiating sites. Proc Natl Acad Sci USA. 1976;73:4070‐4074. doi: 10.1073/PNAS.73.11.4070 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 107. Hanger DP, Anderton BH, Noble W. Tau phosphorylation: the therapeutic challenge for neurodegenerative disease. Trends Mol Med. 2009;15:112‐119. doi: 10.1016/J.MOLMED.2009.01.003 [DOI] [PubMed] [Google Scholar]
- 108. Drepper F, Biernat J, Kaniyappan S, et al. A combinatorial native MS and LC‐MS/MS approach reveals high intrinsic phosphorylation of human Tau but minimal levels of other key modifications. J Biol Chem. 2020;295:18213‐18225. doi: 10.1074/JBC.RA120.015882 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 109. Mair W, Muntel J, Tepper K, et al. FLEXITau: quantifying post‐translational modifications of tau protein in vitro and in human disease. Anal Chem. 2016;88:3704‐3714. doi: 10.1021/ACS.ANALCHEM.5B04509/ASSET/IMAGES/LARGE/AC-2015-04509V_0006.JPEG [DOI] [PMC free article] [PubMed] [Google Scholar]
- 110. Wesseling H, Mair W, Kumar M, et al. Tau PTM profiles identify patient heterogeneity and stages of Alzheimer's disease. Cell. 2020;183:1699‐1713. doi: 10.1016/J.CELL.2020.10.029.e13 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 111. Wegmann S, Biernat J, Mandelkow E. A current view on Tau protein phosphorylation in Alzheimer's disease. Curr Opin Neurobiol. 2021;69:131‐138. doi: 10.1016/J.CONB.2021.03.003 [DOI] [PubMed] [Google Scholar]
- 112. Goedert M, Jakes R. Expression of separate isoforms of human tau protein: correlation with the tau pattern in brain and effects on tubulin polymerization. EMBO J. 1990;9:4225‐4230. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 113. Goode BL, Chau M, Denis PE, Feinstein SC. Structural and functional differences between 3‐repeat and 4‐repeat tau isoforms. Implications for normal tau function and the onset of neurodegenetative disease. J Biol Chem. 2000;275:38182‐38189. doi: 10.1074/jbc.M007489200 [DOI] [PubMed] [Google Scholar]
- 114. Panda D, Samuel JC, Massie M, Feinstein SC, Wilson L. Differential regulation of microtubule dynamics by three‐ and four‐repeat tau: implications for the onset of neurodegenerative disease. Proc Natl Acad Sci. 2003;100:9548‐9553. doi: 10.1073/pnas.1633508100 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 115. Bachmann S, Bell M, Klimek J, Zempel H. Differential effects of the six human TAU isoforms: somatic retention of 2N‐TAU and increased microtubule number induced by 4R‐TAU. Front Neurosci. 2021;15:547. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 116. Rady RM, Zinkowski RP, Binder LI. Presence of tau in isolated nuclei from human brain. Neurobiol Aging. 1995;16:479‐486. doi: 10.1016/0197-4580(95)00023-8 [DOI] [PubMed] [Google Scholar]
- 117. Tashiro K, Hasegawa M, Ihara Y, Iwatsubo T. Somatodendritic localization of phosphorylated tau in neonatal and adult rat cerebral cortex. Neuroreport. 1997;8:2797‐2801. [DOI] [PubMed] [Google Scholar]
- 118. Binder LI, Frankfurter A, Rebhun LI. The distribution of tau in the mammalian central nervous system. J Cell Biol. 1985;101:1371‐1378. doi: 10.1083/JCB.101.4.1371 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 119. Kempf M, Clement A, Faissner A, Lee G, Brandt R. Tau binds to the distal axon early in development of polarity in a microtubule‐ and microfilament‐dependent manner. J Neurosci. 1996. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 120. Paterno G, Bell BM, Gorion KMM, Prokop S, Giasson BI. Reassessment of neuronal tau distribution in adult human brain and implications for tau pathobiology. Acta Neuropathol Commun. 2022;10. doi: 10.1186/S40478-022-01394-9 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 121. Mandell JW, Banker GA. The microtubule cytoskeleton and the development of neuronal polarity. Neurobiol Aging. 1995. doi: 10.1016/0197-4580(94)00164-V [DOI] [PubMed] [Google Scholar]
- 122. Kanaan NM, Grabinski T. Neuronal and glial distribution of tau protein in the adult rat and monkey. Front Mol Neurosci. 2021;14:607303. doi: 10.3389/FNMOL.2021.607303/BIBTEX [DOI] [PMC free article] [PubMed] [Google Scholar]
- 123. Li X, Kumar Y, Zempel H, Mandelkow E, Biernat J, Mandelkow E. Novel diffusion barrier for axonal retention of Tau in neurons and its failure in neurodegeneration. EMBO J. 2011;30:4825‐4837. doi: 10.1038/emboj.2011.376 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 124. Zempel H, Dennissen FJA, Kumar Y, et al. Axodendritic sorting and pathological missorting of Tau are isoform‐specific and determined by axon initial segment architecture. J Biol Chem. 2017;292:12192‐12207. doi: 10.1074/jbc.M117.784702 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 125. Aronov S, Aranda G, Behar L, Ginzburg I. Axonal tau mRNA localization coincides with tau protein in living neuronal cells and depends on axonal targeting signal. J Neurosci. 2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 126. Morita T, Sobuě K. Specification of neuronal polarity regulated by local translation of CRMP2 and tau via the mTOR‐p70S6K pathway. J Biol Chem. 2009. doi: 10.1074/jbc.M109.008177 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 127. Balaji V, Kaniyappan S, Mandelkow E, Wang Y, Mandelkow EM. Pathological missorting of endogenous MAPT/Tau in neurons caused by failure of protein degradation systems. Autophagy. 2018. doi: 10.1080/15548627.2018.1509607 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 128. Hirokawa N, Funakoshi T, Sato‐Harada R, Kanai Y. Selective stabilization of tau in axons and microtubule‐associated protein 2C in cell bodies and dendrites contributes to polarized localization of cytoskeletal proteins in mature neurons. J Cell Biol. 1996. doi: 10.1083/jcb.132.4.667 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 129. Kanai Y, Hirokawa N. Sorting mechanisms of Tau and MAP2 in neurons: suppressed axonal transit of MAP2 and locally regulated microtubule binding. Neuron. 1995. doi: 10.1016/0896-6273(95)90298-8 [DOI] [PubMed] [Google Scholar]
- 130. Scholz T, Mandelkow E. Transport and diffusion of Tau protein in neurons. Cell Mol Life Sci. 2014;71:3139‐3150. doi: 10.1007/S00018-014-1610-7/FIGURES/4 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 131. Zempel H, Mandelkow E. Mechanisms of axonal sorting of tau and influence of the axon initial segment on tau cell polarity. Adv Exp Med Biol. 2019;1184:69‐77. doi: 10.1007/978-981-32-9358-8_6/TABLES/1 [DOI] [PubMed] [Google Scholar]
- 132. Evans DB, Rank KB, Bhattacharya K, Thomsen DR, Gurney ME, Sharma SK. Tau phosphorylation at serine 396 and Serine 404 by human recombinant tau protein kinase II inhibits tau's ability to promote microtubule assembly. J Biol Chem. 2000;275:24977‐24983. doi: 10.1074/jbc.M000808200 [DOI] [PubMed] [Google Scholar]
- 133. Kishi M, Pan YA, Crump JG, Sanes JR. Mammalian SAD kinases are required for neuronal polarization. Science (1979). 2005. doi: 10.1126/science.1107403 [DOI] [PubMed] [Google Scholar]
- 134. Tsushima H, Emanuele M, Polenghi A, et al. HDAC6 and RhoA are novel players in Abeta‐driven disruption of neuronal polarity. Nat Commun. 2015;6:7781. doi: 10.1038/ncomms8781 [DOI] [PubMed] [Google Scholar]
- 135. Ittner A, Chua SW, Bertz J, et al. Site‐specific phosphorylation of tau inhibits amyloid‐β toxicity in Alzheimer's mice. Science (1979). 2016;354:904‐908. doi: 10.1126/SCIENCE.AAH6205/SUPPL_FILE/ITTNER.SM.PDF [DOI] [PubMed] [Google Scholar]
- 136. Fukunaga K, Muller D, Ohmitsu M, Bakó E, DePaoli‐Roach AA, Miyamoto E. Decreased protein phosphatase 2A activity in hippocampal long‐term potentiation. J Neurochem. 2000;74:807‐817. doi: 10.1046/J.1471-4159.2000.740807.X [DOI] [PubMed] [Google Scholar]
- 137. Sontag E, Nunbhakdi‐Craig V, Lee G, Bloom GS, Mumby MC. Regulation of the phosphorylation state and microtubule‐binding activity of Tau by protein phosphatase 2A. Neuron. 1996;17:1201‐1207. doi: 10.1016/S0896-6273(00)80250-0 [DOI] [PubMed] [Google Scholar]
- 138. Buchholz S, Zempel H. Suppression of mature TAU isoforms prevents Alzheimer's disease‐like amyloid‐beta oligomer‐induced spine loss in rodent neurons. Neural Regen Res. 2023;19. doi: accepted manuscript [DOI] [PMC free article] [PubMed] [Google Scholar]
- 139. Liu C, Götz J. Profiling Murine Tau with 0N, 1 and 2N isoform‐specific antibodies in brain and peripheral organs reveals distinct subcellular localization, with the 1N Isoform being enriched in the nucleus. PLoS One. 2014;8:e84849. doi: 10.1371/journal.pone.0084849 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 140. Xia D, Li C, Götz J. Pseudophosphorylation of Tau at distinct epitopes or the presence of the P301L mutation targets the microtubule‐associated protein Tau to dendritic spines. Biochim Biophys Acta Mol Basis Dis. 2015;1852:913‐924. doi: 10.1016/J.BBADIS.2014.12.017 [DOI] [PubMed] [Google Scholar]
- 141. Siano G, Falcicchia C, Origlia N, Cattaneo A, Di Primio C. Non‐canonical roles of tau and their contribution to synaptic dysfunction. Int J Mol Sci. 2021;22:22. doi: 10.3390/IJMS221810145 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 142. Kavanagh T, Halder A, Drummond E. Tau interactome and RNA binding proteins in neurodegenerative diseases. Mol Neurodegener. 2022;17:1‐17. doi: 10.1186/S13024-022-00572-6 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 143. Liu C, Song X, Nisbet R, Götz J. Co‐immunoprecipitation with tau isoform‐specific antibodies reveals distinct protein interactions and highlights a putative role for 2N tau in disease. J Biol Chem. 2016;291:8173‐8188. doi: 10.1074/JBC.M115.641902 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 144. Tracy TE, Madero‐Pérez J, Swaney DL, et al. Tau interactome maps synaptic and mitochondrial processes associated with neurodegeneration. Cell. 2022;185:712‐728. doi: 10.1016/J.CELL.2021.12.041.e14 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 145. Gunawardana CG, Mehrabian M, Wang X, et al. The human tau interactome: binding to the ribonucleoproteome, and impaired binding of the proline‐to‐leucine mutant at position 301 (P301L) to chaperones and the proteasome. Mol Cell Proteomics. 2015;14:3000‐3014. doi: 10.1074/MCP.M115.050724 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 146. Wang X, Williams D, Müller I, et al. Tau interactome analyses in CRISPR‐Cas9 engineered neuronal cells reveal ATPase‐dependent binding of wild‐type but not P301L Tau to non‐muscle myosins. Sci Rep. 2019;9:16238. doi: 10.1038/s41598-019-52543-5 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 147. Meier S, Bell M, Lyons DN, et al. Identification of novel tau interactions with endoplasmic reticulum proteins in Alzheimer's Disease Brain. J Alzheimers Dis. 2015;48:687‐702. doi: 10.3233/JAD-150298 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 148. Hsieh YC, Guo C, Yalamanchili HK, et al. Tau‐mediated disruption of the spliceosome triggers cryptic rna splicing and neurodegeneration in Alzheimer's disease. Cell Rep. 2019;29:301‐316. doi: 10.1016/J.CELREP.2019.08.104/ATTACHMENT/7E4943EB-185C-4FC7-B07D-1E178C8EE1B1/MMC3.XLSX.e10 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 149. Ayyadevara S, Balasubramaniam M, Parcon PA, et al. Proteins that mediate protein aggregation and cytotoxicity distinguish Alzheimer's hippocampus from normal controls. Aging Cell. 2016;15:924‐939. doi: 10.1111/ACEL.12501 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 150. Wu C, Zhao J, Wu Q, Tan Q, Liu Q, Xiao S. Tau N‐terminal inserts regulate tau liquid‐liquid phase separation and condensates maturation in a neuronal cell model. Int J Mol Sci. 2021;22. doi: 10.3390/IJMS22189728 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 151. Zbinden A, Pérez‐Berlanga M, De Rossi P, Polymenidou M. Phase separation and neurodegenerative diseases: a disturbance in the force. Dev Cell. 2020;55:45‐68. doi: 10.1016/J.DEVCEL.2020.09.014 [DOI] [PubMed] [Google Scholar]
- 152. Sinsky J, Majerova P, Kovac A, Kotlyar M, Jurisica I, Hanes J. Physiological tau interactome in brain and its link to tauopathies. J Proteome Res. 2020;19:2429‐2442. doi: 10.1021/ACS.JPROTEOME.0C00137/SUPPL_FILE/PR0C00137_SI_003.XLSX [DOI] [PubMed] [Google Scholar]
- 153. Maziuk BF, Apicco DJ, Cruz AL, et al. RNA binding proteins co‐localize with small tau inclusions in tauopathy. Acta Neuropathol Commun. 2018;6:71. doi: 10.1186/S40478-018-0574-5/TABLES/1 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 154. Yoshiyama Y, Higuchi M, Zhang B, et al. Synapse loss and microglial activation precede tangles in a P301S tauopathy mouse model. Neuron. 2007;53:337‐351. doi: 10.1016/j.neuron.2007.01.010 [DOI] [PubMed] [Google Scholar]
- 155. Wang P, Joberty G, Buist A, et al. Tau interactome mapping based identification of Otub1 as Tau deubiquitinase involved in accumulation of pathological Tau forms in vitro and in vivo. Acta Neuropathol. 2017;133:731‐749. doi: 10.1007/S00401-016-1663-9/FIGURES/8 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 156. Ittner A, Ittner LM. Dendritic tau in Alzheimer's disease. Neuron. 2018;99:13‐27. doi: 10.1016/J.NEURON.2018.06.003 [DOI] [PubMed] [Google Scholar]
- 157. Yu A, Fox SG, Cavallini A, et al. Tau protein aggregates inhibit the protein‐folding and vesicular trafficking arms of the cellular proteostasis network. J Biol Chem. 2019;294:7917‐7930. doi: 10.1074/JBC.RA119.007527 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 158. Blaudin de Thé FX, Lassus B, Schaler AW, et al. P62 accumulates through neuroanatomical circuits in response to tauopathy propagation. Acta Neuropathol Commun. 2021;9. doi: 10.1186/S40478-021-01280-W [DOI] [PMC free article] [PubMed] [Google Scholar]
- 159. Lopes S, Vaz‐Silva J, Pinto V, et al. Tau protein is essential for stress‐induced brain pathology. Proc Natl Acad Sci USA. 2016;113:E3755‐3763. doi: 10.1073/PNAS.1600953113 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 160. Ginsberg SD, Crino PB, Lee VMY, Eberwine JH, Trojanowski JQ. Sequestration of RNA in Alzheimer's disease neurofibrillary tangles and senile plaques. Ann Neurol. 1997;41:200‐209. doi: 10.1002/ANA.410410211 [DOI] [PubMed] [Google Scholar]
- 161. Hsieh YC, Guo C, Yalamanchili HK, et al. Tau‐mediated disruption of the spliceosome triggers cryptic RNA splicing and neurodegeneration in Alzheimer's disease. Cell Rep. 2019;29:301‐316.e10. doi: 10.1016/J.CELREP.2019.08.104 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 162. Evans HT, Taylor D, Kneynsberg A, Bodea LG, Götz J. Altered ribosomal function and protein synthesis caused by tau. Acta Neuropathol Commun. 2021;9. doi: 10.1186/S40478-021-01208-4 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 163. Koren SA, Hamm MJ, Meier SE, et al. Tau drives translational selectivity by interacting with ribosomal proteins. Acta Neuropathol. 2019;137:571‐583. doi: 10.1007/S00401-019-01970-9/FIGURES/7 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 164. Ittner LM, Ke YD, Delerue F, et al. Dendritic function of tau mediates amyloid‐beta toxicity in Alzheimer's disease mouse models. Cell. 2010;142:387‐397. doi: 10.1016/j.cell.2010.06.036 [DOI] [PubMed] [Google Scholar]
- 165. Montalbano M, McAllen S, Sengupta U. Tau oligomers mediate aggregation of RNA‐binding proteins Musashi1 and Musashi2 inducing Lamin alteration. Aging Cell. 2019;18(6):e13035. doi: 10.1111/acel.13035 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 166. Ash PEA, Lei S, Shattuck J, et al. TIA1 potentiates tau phase separation and promotes generation of toxic oligomeric tau. Proc Natl Acad Sci USA. 2021;118:e2014188118. doi: 10.1073/PNAS.2014188118/SUPPL_FILE/PNAS.2014188118.SAPP.PDF [DOI] [PMC free article] [PubMed] [Google Scholar]
- 167. Sultan A, Nesslany F, Violet M, et al. Nuclear tau, a key player in neuronal DNA protection. J Biol Chem. 2011;286:4566‐4575. doi: 10.1074/JBC.M110.199976 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 168. Violet M, Delattre L, Tardivel M, et al. A major role for Tau in neuronal DNA and RNA protection in vivo under physiological and hyperthermic conditions. Front Cell Neurosci. 2014;8:84. doi: 10.3389/fncel.2014.00084 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 169. Kampers T, Friedhoff P, Biernat J, Mandelkow EM, Mandelkow E. RNA stimulates aggregation of microtubule‐associated protein tau into Alzheimer‐like paired helical filaments. FEBS Lett. 1996;399:344‐349. doi: 10.1016/S0014-5793(96)01386-5 [DOI] [PubMed] [Google Scholar]
- 170. Jiang L, Lin W, Zhang C, et al. Interaction of tau with HNRNPA2B1 and N6‐methyladenosine RNA mediates the progression of tauopathy. Mol Cell. 2021;81:4209‐4227. doi: 10.1016/j.molcel.2021.07.038 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 171. Mondragon‐Rodriguez S, Trillaud‐Doppia E, Dudilot A, et al. Interaction of endogenous tau protein with synaptic proteins is regulated by N‐methyl‐D‐aspartate receptor‐dependent tau phosphorylation. J Biol Chem. 2012;287:32040‐32053. doi: 10.1074/jbc.M112.401240 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 172. Regan P, Whitcomb DJ, Cho K. Physiological and pathophysiological implications of synaptic tau. Neuroscientist. 2017;23:137‐151. doi: 10.1177/1073858416633439 [DOI] [PubMed] [Google Scholar]
- 173. Esmaeli‐Azad B, McCarty JH, Feinstein SC. Sense and antisense transfection analysis of tau function: tau influences net microtubule assembly, neurite outgrowth and neuritic stability. J Cell Sci. 1994. [DOI] [PubMed] [Google Scholar]
- 174. Takei Y, Teng J, Harada A, Hirokawa N. Defects in axonal elongation and neuronal migration in mice with disrupted tau and map1b genes. J Cell Biol. 2000;150:989‐1000. doi: 10.1083/JCB.150.5.989 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 175. Sudo H, Baas PW. Strategies for diminishing katanin‐based loss of microtubules in tauopathic neurodegenerative diseases. Hum Mol Genet. 2011;20:763‐778. doi: 10.1093/hmg/ddq521 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 176. Harada A, Oguchi K, Okabe S, et al. Altered microtubule organization in small‐calibre axons of mice lacking tau protein. Nature. 1994;369:488‐491. doi: 10.1038/369488a0 [DOI] [PubMed] [Google Scholar]
- 177. Ma QL, Zuo X, Yang F, et al. Loss of MAP function leads to hippocampal synapse loss and deficits in the Morris Water Maze with aging. J Neurosci. 2014;34:7124‐7136. doi: 10.1523/JNEUROSCI.3439-13.2014 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 178. Lopes S, Lopes A, Pinto V, et al. Absence of Tau triggers age‐dependent sciatic nerve morphofunctional deficits and motor impairment. Aging Cell. 2016;15:208‐216. doi: 10.1111/ACEL.12391 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 179. Sotiropoulos I, Lopes AT, Pinto V, et al. Selective impact of Tau loss on nociceptive primary afferents and pain sensation. Exp Neurol. 2014;261:486‐493. doi: 10.1016/J.EXPNEUROL.2014.07.008 [DOI] [PubMed] [Google Scholar]
- 180. Marciniak E, Leboucher A, Caron E, et al. Tau deletion promotes brain insulin resistance. J Exp Med. 2017;214:2257‐2269. doi: 10.1084/JEM.20161731 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 181. Gonçalves RA, Wijesekara N, Fraser PE, De Felice FG. Behavioral abnormalities in knockout and humanized tau mice. Front Endocrinol. 2020;11:124. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 182. Lei P, Ayton S, Moon S, et al. Motor and cognitive deficits in aged tau knockout mice in two background strains. Mol Neurodegener. 2014;9:29. doi: 10.1186/1750-1326-9-29 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 183. Cantero JL, Hita‐Yaez E, Moreno‐Lopez B, Portillo F, Rubio A, Avila J. Tau protein role in sleep‐wake cycle. J Alzheimers Dis. 2010;21:411‐421. doi: 10.3233/JAD-2010-100285 [DOI] [PubMed] [Google Scholar]
- 184. Ahmed T, Van der Jeugd A, Blum D, et al. Cognition and hippocampal synaptic plasticity in mice with a homozygous tau deletion. Neurobiol Aging. 2014;35:2474‐2478. doi: 10.1016/J.NEUROBIOLAGING.2014.05.005 [DOI] [PubMed] [Google Scholar]
- 185. Criado‐Marrero M, Sabbagh JJ, Jones MR, Chaput D, Dickey CA, Blair LJ. Hippocampal neurogenesis is enhanced in adult tau deficient mice. Cells. 2020;9:210. doi: 10.3390/CELLS9010210 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 186. Caceres A, Kosik KS. Inhibition of neurite polarity by tau antisense oligonucleotides in primary cerebellar neurons. Nature. 1990. doi: 10.1038/343461a0 [DOI] [PubMed] [Google Scholar]
- 187. Caceres A, Potrebic S, Kosik KS. The effect of tau antisense oligonucleotides on neurite formation of cultured cerebellar macroneurons. J Neurosci. 1991;11:1515‐1523. doi: 10.1523/JNEUROSCI.11-06-01515.1991 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 188. Tint I, Slaughter T, Fischer I, Black MM. Acute inactivation of tau has no effect on dynamics of microtubules in growing axons of cultured sympathetic neurons. J Neurosci. 1998;18:8660‐8673. doi: 10.1523/JNEUROSCI.18-21-08660.1998 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 189. Bessone IF, Damianich A, Pedroncini O, et al. Tau conditional reduction in human‐derived neurons exposes novel tau‐associated functions in electrical activity and axonal transport. Alzheimers Dement. 2021;17:e052863. doi: 10.1002/ALZ.052863 [DOI] [Google Scholar]
- 190. Ng B, Vowles J, Bertherat F, et al. Tau depletion in human neurons mitigates Aβ‐driven toxicity. Mol Psychiatry. 2024. doi: 10.1038/s41380-024-02463-2 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 191. Andorfer C, Kress Y, Espinoza M, et al. Hyperphosphorylation and aggregation of tau in mice expressing normal human tau isoforms. J Neurochem. 2003;86:582‐590. doi: 10.1046/J.1471-4159.2003.01879.X [DOI] [PubMed] [Google Scholar]
- 192. Polydoro M, Acker CM, Duff K, Castillo PE, Davies P. Age‐dependent impairment of cognitive and synaptic function in the htau mouse model of tau pathology. J Neurosci. 2009;29:10741‐10749. doi: 10.1523/JNEUROSCI.1065-09.2009 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 193. Komuro Y, Xu G, Bhaskar K, Lamb BT. Human tau expression reduces adult neurogenesis in a mouse model of tauopathy. Neurobiol Aging. 2015;36:2034‐2042. doi: 10.1016/J.NEUROBIOLAGING.2015.03.002 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 194. Cho JD, Kim YA, Rafikian EE, Yang M, Santa‐Maria I. Marked mild cognitive deficits in humanized mouse model of Alzheimer's‐type tau pathology. Front Behav Neurosci. 2021;15:634157. doi: 10.3389/FNBEH.2021.634157/BIBTEX [DOI] [PMC free article] [PubMed] [Google Scholar]
- 195. Saito T, Mihira N, Matsuba Y, et al. Humanization of the entire murine Mapt gene provides a murine model of pathological human tau propagation. J Biol Chem. 2019;294:12754‐12765. doi: 10.1074/jbc.RA119.009487 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 196. Gumucio A, Lannfelt L, Nilsson LNG. Lack of exon 10 in the murine tau gene results in mild sensorimotor defects with aging. BMC Neurosci. 2013;14:1‐13. doi: 10.1186/1471-2202-14-148/TABLES/1 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 197. Zheng M, Jiao L, Tang X, et al. Tau haploinsufficiency causes prenatal loss of dopaminergic neurons in the ventral tegmental area and reduction of transcription factor orthodenticle homeobox 2 expression. FASEB J. 2017;31:3349‐3358. doi: 10.1096/FJ.201601303R [DOI] [PMC free article] [PubMed] [Google Scholar]
- 198. Zempel H, The MAPT Isoform 0N3R Is Essential for Human Brain Development and Intolerant to Haploinsufficiency—Loss of Function as Disease Mechanism for TAU Associated Disease 2023. doi: 10.20944/PREPRINTS202312.0513.V1 [DOI]
- 199. Sennvik K, Boekhoorn K, Lasrado R, et al. Tau‐4R suppresses proliferation and promotes neuronal differentiation in the hippocampus of tau knockin/knockout mice. FASEB J. 2007;21:2149‐2161. doi: 10.1096/FJ.06-7735COM [DOI] [PubMed] [Google Scholar]
- 200. Terwel D, Lasrado R, Snauwaert J, et al. Changed conformation of mutant Tau‐P301L underlies the moribund tauopathy, absent in progressive, nonlethal axonopathy of Tau‐4R/2N Transgenic Mice. J Biol Chem. 2005;280:3963‐3973. doi: 10.1074/JBC.M409876200 [DOI] [PubMed] [Google Scholar]
- 201. Liazoghli D, Perreault S, Micheva KD, Desjardins M, Leclerc N. Fragmentation of the Golgi apparatus induced by the overexpression of wild‐type and mutant human tau forms in neurons. Am J Pathol. 2005;166:1499‐1514. doi: 10.1016/S0002-9440(10)62366-8 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 202. Sola M, Magrin C, Pedrioli G, et al. Tau affects P53 function and cell fate during the DNA damage response. Commun Biol. 2020;3:245. doi: 10.1038/s42003-020-0975-4 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 203. Zempel H. Genetic and sporadic forms of tauopathies–TAU as a disease driver for the majority of patients but the minority of tauopathies. Cytoskeleton. 2024;81:66‐70. doi: 10.1002/CM.21793 [DOI] [PubMed] [Google Scholar]
- 204. Koolen DA, Vissers LELM, Pfundt R, et al. A new chromosome 17q21.31 microdeletion syndrome associated with a common inversion polymorphism. Nat Genet. 2006;38:999‐1001. doi: 10.1038/ng1853 [DOI] [PubMed] [Google Scholar]
- 205. Shaw‐Smith C, Pittman AM, Willatt L, et al. Microdeletion encompassing MAPT at chromosome 17q21.3 is associated with developmental delay and learning disability. Nat Genet. 2006;38:1032‐1037. doi: 10.1038/ng1858 [DOI] [PubMed] [Google Scholar]
- 206. Koolen DA, Pfundt R, Linda K, et al. The Koolen‐de Vries syndrome: a phenotypic comparison of patients with a 17q21.31 microdeletion versus a KANSL1 sequence variant. Eur J Hum Genet. 2015;24:652‐659. doi: 10.1038/ejhg.2015.178 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 207. Götz J, Halliday G, Nisbet RM. Molecular pathogenesis of the tauopathies. Annu Rev Pathol. 2019;14:239‐261. doi: 10.1146/annurev-pathmechdis-012418-012936 [DOI] [PubMed] [Google Scholar]
- 208. Shi Y, Zhang W, Yang Y, et al. Structure‐based classification of tauopathies. Nature. 2021;598:359‐363. doi: 10.1038/s41586-021-03911-7 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 209. Kovacs GG, Ghetti B, Goedert M. Classification of diseases with accumulation of Tau protein. Neuropathol Appl Neurobiol. 2022;48:e12792. doi: 10.1111/NAN.12792 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 210. Panza F, Lozupone M, Seripa D, et al. Development of disease‐modifying drugs for frontotemporal dementia spectrum disorders. Nat Rev Neurol. 2020;16:213‐228. doi: 10.1038/s41582-020-0330-x [DOI] [PubMed] [Google Scholar]
- 211. Guo T, Noble W, Hanger DP. Roles of tau protein in health and disease. Acta Neuropathol. 2017;133:665‐704. doi: 10.1007/S00401-017-1707-9 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 212. Catarina Silva M, Haggarty SJ. Tauopathies: deciphering disease mechanisms to develop effective therapies. Int J Mol Sci. 2020;21:8948. doi: 10.3390/IJMS21238948 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 213. D'Souza I, Poorkaj P, Hong M, et al. Missense and silent tau gene mutations cause frontotemporal dementia with parkinsonism‐chromosome 17 type, by affecting multiple alternative RNA splicing regulatory elements. Proc Natl Acad Sci. 1999;96:5598‐5603. doi: 10.1073/PNAS.96.10.5598 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 214. Lee Virginia MY, Zhukareva V, Vogelsberg‐Ragaglia V, et al. Mutation‐specific functional impairments in distinct tau isoforms of hereditary FTDP‐17. Science (1979). 1998;282:1914‐1917. doi: 10.1126/SCIENCE.282.5395.1914 [DOI] [PubMed] [Google Scholar]
- 215. Von Bergen M, Barghorn S, Li L, et al. Mutations of tau protein in frontotemporal dementia promote aggregation of paired helical filaments by enhancing local β‐structure. J Biol Chem. 2001;276:48165‐48174. doi: 10.1074/JBC.M105196200 [DOI] [PubMed] [Google Scholar]
- 216. Ishigaki S, Fujioka Y, Okada Y, et al. Altered tau isoform ratio caused by loss of FUS and SFPQ function leads to FTLD‐like phenotypes. Cell Rep. 2017;18:1118‐1131. doi: 10.1016/J.CELREP.2017.01.013 [DOI] [PubMed] [Google Scholar]
- 217. Orozco D, Tahirovic S, Rentzsch K, Schwenk BM, Haass C, Edbauer D. Loss of fused in sarcoma (FUS) promotes pathological Tau splicing. EMBO Rep. 2012;13:759‐764. doi: 10.1038/EMBOR.2012.90 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 218. Dinkel PD, Siddiqua A, Huynh H, Shah M, Margittai M. Variations in filament conformation dictate seeding barrier between three‐ and four‐repeat tau. Biochemistry. 2011;50:4330‐4336. doi: 10.1021/BI2004685 [DOI] [PubMed] [Google Scholar]
- 219. Combs B, Voss K, Gamblin TC. Pseudohyperphosphorylation has differential effects on polymerization and function of tau isoforms. Biochemistry. 2011;50:9446‐9456. doi: 10.1021/bi2010569 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 220. Adams SJ, de Ture MA, McBride M, Dickson DW, Petrucelli L. Three repeat isoforms of tau inhibit assembly of four repeat tau filaments. PLoS One. 2010;5. doi: 10.1371/JOURNAL.PONE.0010810 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 221. Chakraborty P, Rivière G, Hebestreit A, et al. Acetylation discriminates disease‐specific tau deposition. Nat Commun. 2023;14:1‐13. doi: 10.1038/s41467-023-41672-1 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 222. Buee L, Delacourte A. Comparative biochemistry of tau in progressive supranuclear palsy, corticobasal degeneration, FTDP‐17 and Pick's disease. Brain Pathol. 1999;9:681‐693. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 223. Zhong Q, Congdon EE, Nagaraja HN, Kuret J. Tau isoform composition influences rate and extent of filament formation. J Biol Chem. 2012;287:20711‐20719. doi: 10.1074/JBC.M112.364067 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 224. Losev Y, Paul A, Frenkel‐Pinter M, et al. Novel model of secreted human tau protein reveals the impact of the abnormal N‐glycosylation of tau on its aggregation propensity. Sci Rep. 2019;9. doi: 10.1038/S41598-019-39218-X [DOI] [PMC free article] [PubMed] [Google Scholar]
- 225. Lefebvre T, Ferreira S, Dupont‐Wallois L, et al. Evidence of a balance between phosphorylation and O‐GlcNAc glycosylation of Tau proteins—A role in nuclear localization. Biochim Biophys Acta. 2003;1619:167‐176. doi: 10.1016/S0304-4165(02)00477-4 [DOI] [PubMed] [Google Scholar]
- 226. Brook JD, McCurrach ME, Harley HG, et al. Molecular basis of myotonic dystrophy: expansion of a trinucleotide (CTG) repeat at the 3′ end of a transcript encoding a protein kinase family member. Cell. 1992;68:799‐808. doi: 10.1016/0092-8674(92)90154-5 [DOI] [PubMed] [Google Scholar]
- 227. Liu J, Guo ZN, Yan XL, Yang Y, Huang S. Brain pathogenesis and potential therapeutic strategies in myotonic dystrophy type 1. Front Aging Neurosci. 2021;13:755392. doi: 10.3389/FNAGI.2021.755392/BIBTEX [DOI] [PMC free article] [PubMed] [Google Scholar]
- 228. Dhaenens CM, Schraen‐Maschke S, Tran H, et al. Overexpression of MBNL1 fetal isoforms and modified splicing of Tau in the DM1 brain: two individual consequences of CUG trinucleotide repeats. Exp Neurol. 2008;210:467‐478. doi: 10.1016/J.EXPNEUROL.2007.11.020 [DOI] [PubMed] [Google Scholar]
- 229. Dhaenens CM, Tran H, Frandemiche ML, et al. Mis‐splicing of Tau exon 10 in myotonic dystrophy type 1 is reproduced by overexpression of CELF2 but not by MBNL1 silencing. Biochim Biophys Acta Mol Basis Dis. 2011;1812:732‐742. doi: 10.1016/J.BBADIS.2011.03.010 [DOI] [PubMed] [Google Scholar]
- 230. Weijs R, Okkersen K, van Engelen B, et al. Human brain pathology in myotonic dystrophy type 1: a systematic review. Neuropathology. 2021;41:3‐20. doi: 10.1111/NEUP.12721 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 231. Caillet‐Boudin ML, Fernandez‐Gomez FJ, Tran H, Dhaenens CM, Buee L, Sergeant N. Brain pathology in myotonic dystrophy: when tauopathy meets spliceopathy and RNAopathy. Front Mol Neurosci. 2014;6:73151. doi: 10.3389/FNMOL.2013.00057/BIBTEX [DOI] [PMC free article] [PubMed] [Google Scholar]
- 232. Jiang H, Mankodi A, Swanson MS, Moxley RT, Thornton CA. Myotonic dystrophy type 1 is associated with nuclear foci of mutant RNA, sequestration of muscleblind proteins and deregulated alternative splicing in neurons. Hum Mol Genet. 2004;13:3079‐3088. doi: 10.1093/HMG/DDH327 [DOI] [PubMed] [Google Scholar]
- 233. Carpentier C, Ghanem D, Fernandez‐Gomez FJ, et al. Tau exon 2 responsive elements deregulated in myotonic dystrophy type I are proximal to exon 2 and synergistically regulated by MBNL1 and MBNL2. Biochim Biophys Acta. 2014;1842:654‐664. doi: 10.1016/J.BBADIS.2014.01.004 [DOI] [PubMed] [Google Scholar]
- 234. Rex CS, Gavin CF, Rubio MD, et al. Myosin IIb regulates actin dynamics during synaptic plasticity and memory formation. Neuron. 2010;67:603‐617. doi: 10.1016/J.NEURON.2010.07.016 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 235. Hodges JL, Newell‐Litwa K, Asmussen H, Vicente‐Manzanares M, Horwitz AR. Myosin IIB activity and phosphorylation status determines dendritic spine and post‐synaptic density morphology. PLoS One. 2011;6:e24149. doi: 10.1371/JOURNAL.PONE.0024149 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 236. Rubio MD, Johnson R, Miller CA, Huganir RL, Rumbaugh G. Regulation of synapse structure and function by distinct myosin II motors. J Neurosci. 2011;31:1448‐1460. doi: 10.1523/JNEUROSCI.3294-10.2011 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 237. Parra BC, Maria GA, Madero PJ, et al. Human iPSC 4R tauopathy model uncovers modifiers of tau propagation. BioRxiv 2023:2023.06.19.544278. doi: 10.1101/2023.06.19.544278 [DOI]
- 238. Sealey MA, Vourkou E, Cowan CM, et al. Distinct phenotypes of three‐repeat and four‐repeat human tau in a transgenic model of tauopathy. Neurobiol Dis. 2017;105:74‐83. doi: 10.1016/J.NBD.2017.05.003 [DOI] [PubMed] [Google Scholar]
- 239. 2023 Alzheimer's disease facts and figures. Alzheimers Dement. 2023;19:1598‐1695. doi: 10.1002/ALZ.13016 [DOI] [PubMed] [Google Scholar]
- 240. Wilson DM, Cookson MR, Van Den Bosch L, Zetterberg H, Holtzman DM, Dewachter I. Hallmarks of neurodegenerative diseases. Cell. 2023;186:693‐714. doi: 10.1016/J.CELL.2022.12.032 [DOI] [PubMed] [Google Scholar]
- 241. Cacace R, Sleegers K, Van Broeckhoven C. Molecular genetics of early‐onset Alzheimer's disease revisited. Alzheimers Dement. 2016;12:733‐748. doi: 10.1016/J.JALZ.2016.01.012 [DOI] [PubMed] [Google Scholar]
- 242. Guo T, Zhang D, Zeng Y, Huang TY, Xu H, Zhao Y. Molecular and cellular mechanisms underlying the pathogenesis of Alzheimer's disease. Mol Neurodegener. 2020;15:1‐37. doi: 10.1186/S13024-020-00391-7/FIGURES/2 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 243. Haass C, Kaether C, Thinakaran G, Sisodia S. Trafficking and proteolytic processing of APP. Cold spring harb. Perspect Med. 2012;2:a006270. doi: 10.1101/CSHPERSPECT.A006270 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 244. Thinakaran G, Koo EH. Amyloid precursor protein trafficking, processing, and function. J Biol Chem. 2008;283:29615‐29619. doi: 10.1074/JBC.R800019200 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 245. Galanis C, Fellenz M, Becker D, et al. Amyloid‐beta mediates homeostatic synaptic plasticity. J Neurosci. 2021;41:5157‐5172. doi: 10.1523/JNEUROSCI.1820-20.2021 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 246. Puzzo D, Privitera L, Leznik E, et al. Picomolar amyloid‐β positively modulates synaptic plasticity and memory in hippocampus. J Neurosci. 2008;28:14537‐14545. doi: 10.1523/JNEUROSCI.2692-08.2008 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 247. Wei W, Nguyen LN, Kessels HW, Hagiwara H, Sisodia S, Malinow R. Amyloid beta from axons and dendrites reduces local spine number and plasticity. Nat Neurosci. 2009;13:190‐196. doi: 10.1038/nn.2476 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 248. Kamenetz F, Tomita T, Hsieh H, et al. APP processing and synaptic function. Neuron. 2003;37:925‐937. doi: 10.1016/S0896-6273(03)00124-7 [DOI] [PubMed] [Google Scholar]
- 249. Cirrito JR, Yamada KA, Finn MB, et al. Synaptic activity regulates interstitial fluid amyloid‐β levels in vivo. Neuron. 2005;48:913‐922. doi: 10.1016/J.NEURON.2005.10.028 [DOI] [PubMed] [Google Scholar]
- 250. Zhou B, Lu JG, Siddu A, Wernig M, Südhof TC. Synaptogenic effect of APP‐Swedish mutation in familial Alzheimer's disease. Sci Transl Med. 2022;14. doi: 10.1126/SCITRANSLMED.ABN9380/SUPPL_FILE/SCITRANSLMED.ABN9380_TABLE_S1.ZIP [DOI] [PMC free article] [PubMed] [Google Scholar]
- 251. Benarroch EE. Glutamatergic synaptic plasticity and dysfunction in Alzheimer disease. Neurology. 2018;91:125‐132. doi: 10.1212/WNL.0000000000005807 [DOI] [PubMed] [Google Scholar]
- 252. Selkoe DJ, Hardy J. The amyloid hypothesis of Alzheimer's disease at 25 years. EMBO Mol Med. 2016;8:595‐608. doi: 10.15252/EMMM.201606210 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 253. Spires‐Jones TL, Hyman BT. The intersection of amyloid beta and tau at synapses in Alzheimer's disease. Neuron. 2014;82:756‐771. doi: 10.1016/j.neuron.2014.05.004 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 254. Kelleher RJ, Shen J. Presenilin‐1 mutations and Alzheimer's disease. Proc Natl Acad Sci USA. 2017;114:629. doi: 10.1073/PNAS.1619574114 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 255. Dong L, Liu C, Sha L, et al. PSEN2 mutation spectrum and novel functionally validated mutations in Alzheimer's disease: data from PUMCH dementia cohort. J Alzheimers Dis. 2022;87:1549. doi: 10.3233/JAD-220194 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 256. van der Lee SJ, Wolters FJ, Ikram MK, et al. The effect of APOE and other common genetic variants on the onset of Alzheimer's disease and dementia: a community‐based cohort study. Lancet Neurol. 2018;17:434‐444. doi: 10.1016/S1474-4422(18)30053-X [DOI] [PubMed] [Google Scholar]
- 257. Efthymiou AG, Goate AM. Late onset Alzheimer's disease genetics implicates microglial pathways in disease risk. Mol Neurodegener. 2017;12:1‐12. doi: 10.1186/S13024-017-0184-X/FIGURES/2 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 258. Jansen IE, Savage JE, Watanabe K, et al. Genome‐wide meta‐analysis identifies new loci and functional pathways influencing Alzheimer's disease risk. Nat Genet. 2019;51:404‐413. doi: 10.1038/s41588-018-0311-9 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 259. Kunkle BW, Grenier‐Boley B, Sims R, et al. Genetic meta‐analysis of diagnosed Alzheimer's disease identifies new risk loci and implicates Aβ, tau, immunity and lipid processing. Nat Genet. 2019;51:414‐430. doi: 10.1038/S41588-019-0358-2 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 260. Zhang B, Gaiteri C, Bodea LG, et al. Integrated systems approach identifies genetic nodes and networks in late‐onset Alzheimer's disease. Cell. 2013;153:707‐720. doi: 10.1016/J.CELL.2013.03.030 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 261. Reiman EM, Arboleda‐Velasquez JF, Quiroz YT, et al. Exceptionally low likelihood of Alzheimer's dementia in APOE2 homozygotes from a 5,000‐person neuropathological study. Nat Commun. 2020;11:1‐11. doi: 10.1038/s41467-019-14279-8 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 262. Genin E, Hannequin D, Wallon D, et al. APOE and Alzheimer disease: a major gene with semi‐dominant inheritance. Mol Psychiatry. 2011;16:903‐907. doi: 10.1038/MP.2011.52 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 263. Jonsson T, Atwal JK, Steinberg S, et al. A mutation in APP protects against Alzheimer's disease and age‐related cognitive decline. Nature. 2012;488:96‐99. doi: 10.1038/nature11283 [DOI] [PubMed] [Google Scholar]
- 264. Arboleda‐Velasquez JF, Lopera F, O'Hare M, et al. Resistance to autosomal dominant Alzheimer's disease in an APOE3 Christchurch homozygote: a case report. Nat Med. 2019;25:1680‐1683. doi: 10.1038/s41591-019-0611-3 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 265. Pontecorvo MJ, et al. investigators for the 18F‐A‐1451‐A, Devous MD, investigators for the 18F‐A‐1451‐A, Kennedy I, investigators for the 18F‐A‐1451‐A, et al. A multicentre longitudinal study of flortaucipir (18F) in normal ageing, mild cognitive impairment and Alzheimer's disease dementia. Brain. 2019;142:1723‐1735. doi: 10.1093/BRAIN/AWZ090 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 266. Price JL, McKeel DW, Buckles VD, et al. Neuropathology of nondemented aging: presumptive evidence for preclinical Alzheimer disease. Neurobiol Aging. 2009;30:1026‐1036. doi: 10.1016/J.NEUROBIOLAGING.2009.04.002 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 267. Price JL, Morris JC, Tangles and Plaques in Nondemented Aging and “Preclinical” Alzheimer's Disease 1999. doi: 10.1002/1531-8249 [DOI] [PubMed]
- 268. Giannakopoulos P, Herrmann FR, Bussière T, et al. Tangle and neuron numbers, but not amyloid load, predict cognitive status in Alzheimer's disease. Neurology. 2003;60:1495‐1500. doi: 10.1212/01.WNL.0000063311.58879.01 [DOI] [PubMed] [Google Scholar]
- 269. Nelson PT, Alafuzoff I, Bigio EH, et al. Correlation of Alzheimer disease neuropathologic changes with cognitive status: a review of the literature. J Neuropathol Exp Neurol. 2012;71:362‐381. doi: 10.1097/NEN.0B013E31825018F7 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 270. Roberson ED, Scearce‐Levie K, Palop JJ, et al. Reducing endogenous tau ameliorates amyloid beta‐induced deficits in an Alzheimer's disease mouse model. Science. 2007;316:750‐754. doi: 10.1126/SCIENCE.1141736 [DOI] [PubMed] [Google Scholar]
- 271. Ke YD, Suchowerska AK, Van Der Hoven J, et al. Lessons from Tau‐deficient mice. Int J Alzheimers Dis. 2012. doi: 10.1155/2012/873270 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 272. Vossel KA, Zhang K, Brodbeck J, et al. Tau reduction prevents Abeta‐induced defects in axonal transport. Science. 2010;330. doi: 10.1126/SCIENCE.1194653 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 273. Jagust W. Is amyloid‐β harmful to the brain? Insights from human imaging studies. Brain. 2016;139:23‐30. doi: 10.1093/BRAIN/AWV326 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 274. Desikan RS, McEvoy LK, Thompson WK, Holland D, Rddey JC, Blennow K, et al. Amyloid‐β associated volume loss occurs only in the presence of phospho‐tau. Ann Neurol. 2011;70:657‐661. doi: 10.1002/ANA.22509 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 275. Crystal H, Dickson D, Fuld P, et al. Clinico‐pathologic studies in dementia: nondemented subjects with pathologically confirmed Alzheimer's disease. Neurology. 1988;38:1682‐1687. doi: 10.1212/WNL.38.11.1682 [DOI] [PubMed] [Google Scholar]
- 276. Shipton OA, Leitz JR, Dworzak J, et al. Tau protein is required for amyloid {beta}‐induced impairment of hippocampal long‐term potentiation. J Neurosci. 2011;31:1688‐1692. doi: 10.1523/jneurosci.2610-10.2011 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 277. DeVos SL, Miller RL, Schoch KM, et al. Tau reduction prevents neuronal loss and reverses pathological tau deposition and seeding in mice with tauopathy. Sci Transl Med. 2017;9. doi: 10.1126/scitranslmed.aag0481 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 278. Chin J, Palop JJ, Yu GQ, Kojima N, Masliah E, Mucke L. Fyn kinase modulates synaptotoxicity, but not aberrant sprouting, in human amyloid precursor protein transgenic mice. J Neurosci. 2004;24:4692‐4697. doi: 10.1523/JNEUROSCI.0277-04.2004 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 279. Roberson ED, Halabisky B, Yoo JW, et al. Amyloid‐β/Fyn‐induced synaptic, network, and cognitive impairments depend on tau levels in multiple mouse models of Alzheimer's disease. J Neurosci. 2011;31:700. doi: 10.1523/JNEUROSCI.4152-10.2011 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 280. Morris M, Koyama A, Masliah E, Mucke L. Tau reduction does not prevent motor deficits in two mouse models of Parkinson's disease. PLoS One. 2011;6. doi: 10.1371/JOURNAL.PONE.0029257 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 281. Pacheco CD, Elrick MJ, Lieberman AP. Tau deletion exacerbates the phenotype of Niemann‐Pick type C mice and implicates autophagy in pathogenesis. Hum Mol Genet. 2009;18:956‐965. doi: 10.1093/HMG/DDN423 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 282. Brettschneider J, Del Tredici K, Lee VMY, Trojanowski JQ. Spreading of pathology in neurodegenerative diseases: a focus on human studies. Nat Rev Neurosci. 2015;16:109‐120. doi: 10.1038/nrn3887 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 283. Delacourte A, David JP, Sergeant N, et al. The biochemical pathway of neurofibrillary degeneration in aging and Alzheimer's disease. Neurology. 1999;52:1158‐1165. doi: 10.1212/WNL.52.6.1158 [DOI] [PubMed] [Google Scholar]
- 284. Braak H, Del Tredici K. Spreading of tau pathology in sporadic Alzheimer's disease along cortico‐cortical top‐down connections. Cereb Cortex. 2018;28:3372‐3384. doi: 10.1093/cercor/bhy152 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 285. Iba M, Guo JL, McBride JD, Zhang B, Trojanowski JQ, Lee VMY. Synthetic tau fibrils mediate transmission of neurofibrillary tangles in a transgenic mouse model of Alzheimer's‐like tauopathy. J Neurosci. 2013;33:1024‐1037. doi: 10.1523/JNEUROSCI.2642-12.2013 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 286. Liu L, Drouet V, Wu JW, et al. Trans‐synaptic spread of tau pathology in vivo. PLoS One. 2012;7:e31302. doi: 10.1371/JOURNAL.PONE.0031302 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 287. Clavaguera F, Bolmont T, Crowther RA, et al. Transmission and spreading of tauopathy in transgenic mouse brain. Nat Cell Biol. 2009;11:909‐913. doi: 10.1038/ncb1901 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 288. Clavaguera F, Akatsu H, Fraser G, et al. Brain homogenates from human tauopathies induce tau inclusions in mouse brain. Proc Natl Acad Sci USA. 2013;110:9535‐9540. doi: 10.1073/PNAS.1301175110/SUPPL_FILE/SD01.PDF [DOI] [PMC free article] [PubMed] [Google Scholar]
- 289. De Calignon A, Polydoro M, Suárez‐Calvet M, et al. Propagation of tau pathology in a model of early Alzheimer's disease. Neuron. 2012;73:685‐697. doi: 10.1016/J.NEURON.2011.11.033 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 290. DeVos SL, Corjuc BT, Oakley DH, et al. Synaptic tau seeding precedes tau pathology in human Alzheimer's disease brain. Front Neurosci. 2018;12:267. doi: 10.3389/FNINS.2018.00267/BIBTEX [DOI] [PMC free article] [PubMed] [Google Scholar]
- 291. Aguzzi A, Rajendran L. The transcellular spread of cytosolic amyloids, prions, and prionoids. Neuron. 2009;64:783‐790. doi: 10.1016/j.neuron.2009.12.016 [DOI] [PubMed] [Google Scholar]
- 292. Ganguly P, Do TD, Larini L, et al. Tau assembly: the dominant role of PHF6 (VQIVYK) in microtubule binding region repeat R3. J Phys Chem B. 2015;119:4582‐4593. doi: 10.1021/ACS.JPCB.5B00175 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 293. Xia Y, Xia Y, Prokop S, et al. Tau Ser208 phosphorylation promotes aggregation and reveals neuropathologic diversity in Alzheimer's disease and other tauopathies. Acta Neuropathol Commun. 2020;8:1‐17. doi: 10.1186/S40478-020-00967-W/FIGURES/8 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 294. Despres C, Byrne C, Qi H, et al. Identification of the Tau phosphorylation pattern that drives its aggregation. Proc Natl Acad Sci USA. 2017;114:9080‐9085. doi: 10.1073/PNAS.1708448114/SUPPL_FILE/PNAS.1708448114.SAPP.PDF [DOI] [PMC free article] [PubMed] [Google Scholar]
- 295. Augustinack JC, Schneider A, Mandelkow EM, Hyman BT. Specific tau phosphorylation sites correlate with severity of neuronal cytopathology in Alzheimer's disease. Acta Neuropathol. 2002;103:26‐35. doi: 10.1007/S004010100423/METRICS [DOI] [PubMed] [Google Scholar]
- 296. Dujardin S, Bégard S, Caillierez R, et al. Different tau species lead to heterogeneous tau pathology propagation and misfolding. Acta Neuropathol Commun. 2018;6:132. doi: 10.1186/s40478-018-0637-7 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 297. Kuret J, Chirita CN, Congdon EE, et al. Pathways of tau fibrillization. Biochim Biophys Acta Mol Basis Dis. 2005;1739:167‐178. doi: 10.1016/J.BBADIS.2004.06.016 [DOI] [PubMed] [Google Scholar]
- 298. Zempel H, Mandelkow EM. Linking amyloid‐beta and tau: amyloid‐beta induced synaptic dysfunction via local wreckage of the neuronal cytoskeleton. Neurodegener Dis. 2012;10:64‐72. doi: 10.1159/000332816 [DOI] [PubMed] [Google Scholar]
- 299. Zempel H, Mandelkow E. Lost after translation: missorting of Tau protein and consequences for Alzheimer disease. Trends Neurosci. 2014;37:721‐732. doi: 10.1016/J.TINS.2014.08.004 [DOI] [PubMed] [Google Scholar]
- 300. Hoover BR, Reed MN, Su J, et al. Tau mislocalization to dendritic spines mediates synaptic dysfunction independently of neurodegeneration. Neuron. 2010;68:1067‐1081. doi: 10.1016/j.neuron.2010.11.030 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 301. Dixit R, Ross JL, Goldman YE, Holzbaur ELF. Differential regulation of dynein and kinesin motor proteins by tau. Science. 2008;319:1086‐1089. doi: 10.1126/SCIENCE.1152993 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 302. Fernandez‐Valenzuela JJ, Sanchez‐Varo R, Muñoz‐Castro C, et al. Enhancing microtubule stabilization rescues cognitive deficits and ameliorates pathological phenotype in an amyloidogenic Alzheimer's disease model. Sci Rep. 2020;10:1‐17. doi: 10.1038/s41598-020-71767-4 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 303. Zhang B, Carroll J, Trojanowski JQ, et al. The microtubule‐stabilizing agent, epothilone d, reduces axonal dysfunction, neurotoxicity, cognitive deficits, and Alzheimer‐like pathology in an interventional study with aged tau transgenic mice. J Neurosci. 2012;32:3601‐3611. doi: 10.1523/JNEUROSCI.4922-11.2012 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 304. Brunden KR, Yao Y, Potuzak JS, et al. The characterization of microtubule‐stabilizing drugs as possible therapeutic agents for Alzheimer's disease and related tauopathies. Pharmacol Res. 2011;63:341‐351. doi: 10.1016/J.PHRS.2010.12.002 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 305. Ballatore C, Brunden KR, Huryn DM, Trojanowski JQ, Lee VMY, Smith AB. Microtubule stabilizing agents as potential treatment for alzheimers disease and related neurodegenerative tauopathies. J Med Chem. 2012;55:8979‐8996. doi: 10.1021/JM301079Z/ASSET/IMAGES/LARGE/JM-2012-01079Z_0016.JPEG [DOI] [PMC free article] [PubMed] [Google Scholar]
- 306. Vourkou E, Paspaliaris V, Bourouliti A, et al. Differential effects of human tau isoforms to neuronal dysfunction and toxicity in the drosophila CNS. Int J Mol Sci. 2022;23:12985. doi: 10.3390/IJMS232112985/S1 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 307. Evans HT, Benetatos J, van Roijen M, Bodea L, Götz J. Decreased synthesis of ribosomal proteins in tauopathy revealed by non‐canonical amino acid labelling. EMBO J. 2019;38. doi: 10.15252/EMBJ.2018101174 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 308. Meier S, Bell M, Lyons DN, et al. Pathological tau promotes neuronal damage by impairing ribosomal function and decreasing protein synthesis. J Neurosci. 2016;36:957‐962. doi: 10.1523/JNEUROSCI.3029-15.2016 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 309. Perbet R, Zufferey V, Leroux E, et al. Tau transfer via extracellular vesicles disturbs the astrocytic mitochondrial system. Cells. 2023;12:985. doi: 10.3390/CELLS12070985/S1 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 310. Brunello CA, Merezhko M, Uronen RL, Huttunen HJ. Mechanisms of secretion and spreading of pathological tau protein. Cell Mol Life Sci. 2019;77:1721‐1744. doi: 10.1007/S00018-019-03349-1 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 311. Cummings J, Zhou Y, Lee G, Zhong K, Fonseca J, Cheng F. Alzheimer's disease drug development pipeline: 2023. Alzheimers Dement. 2023;9:e12385. doi: 10.1002/TRC2.12385 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 312. Mummery CJ, Börjesson‐Hanson A, Blackburn DJ, et al. Tau‐targeting antisense oligonucleotide MAPTRx in mild Alzheimer's disease: a phase 1b, randomized, placebo‐controlled trial. Nat Med. 2023;29:1437‐1447. doi: 10.1038/s41591-023-02326-3 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 313. DeVos SL, Goncharoff DK, Chen G, et al. Antisense reduction of tau in adult mice protects against seizures. J Neurosci. 2013;33:12887‐12897. doi: 10.1523/JNEUROSCI.2107-13.2013 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 314. Wegmann S, DeVos SL, Zeitler B, et al. Persistent repression of tau in the brain using engineered zinc finger protein transcription factors. Sci Adv. 2021;7. doi: 10.1126/SCIADV.ABE1611 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 315. Self WK, Holtzman DM. Emerging diagnostics and therapeutics for Alzheimer disease. Nat Med. 2023;2023:1‐13. doi: 10.1038/s41591-023-02505-2 [DOI] [PubMed] [Google Scholar]
- 316. Imbimbo BP, Ippati S, Watling M, Balducci C. A critical appraisal of tau‐targeting therapies for primary and secondary tauopathies. Alzheimers Dement. 2022;18:1008‐1037. doi: 10.1002/ALZ.12453 [DOI] [PubMed] [Google Scholar]
- 317. Riedel G, Klein J, Niewiadomska G, et al. Mechanisms of anticholinesterase interference with tau aggregation inhibitor activity in a tau‐transgenic mouse model. Curr Alzheimer Res. 2020;17:285‐296. doi: 10.2174/1567205017666200224120926 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 318. Gauthier S, Feldman HH, Schneider LS, et al. Efficacy and safety of tau‐aggregation inhibitor therapy in patients with mild or moderate Alzheimer's disease: a randomised, controlled, double‐blind, parallel‐arm, phase 3 trial. Lancet. 2016;388:2873‐2884. doi: 10.1016/S0140-6736(16)31275-2 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 319. Wilcock GK, Gauthier S, Frisoni GB, et al. Potential of low dose leuco‐Methylthioninium Bis(Hydromethanesulphonate) (LMTM) monotherapy for treatment of mild Alzheimer's disease: cohort analysis as modified primary outcome in a phase III clinical trial. J Alzheimers Dis. 2018;61:435‐457. doi: 10.3233/JAD-170560 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 320.ACI‐3024 | ALZFORUM n.d. https://www.alzforum.org/therapeutics/aci‐3024 (accessed December 6, 2023)
- 321. Lovestone S, Davis DR, Webster MT, et al. Lithium reduces tau phosphorylation: effects in living cells and in neurons at therapeutic concentrations. Biol Psychiatry. 1999;45. doi: 10.1016/S0006-3223(98)00183-8 [DOI] [PubMed] [Google Scholar]
- 322. Long ZM, Zhao L, Jiang R, et al. Valproic acid modifies synaptic structure and accelerates neurite outgrowth via the glycogen synthase kinase‐3β signaling pathway in an Alzheimer's disease model. CNS Neurosci Ther. 2015;21:887‐897. doi: 10.1111/CNS.12445 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 323. Liu F, Shi J, Tanimukai H, et al. Reduced O‐GlcNAcylation links lower brain glucose metabolism and tau pathology in Alzheimer's disease. Brain. 2009;132:1820‐1832. doi: 10.1093/BRAIN/AWP099 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 324. Liu F, Iqbal K, Grundke‐Iqbal I, Hart GW. Gong CX. O‐GlcNAcylation regulates phosphorylation of tau: a mechanism involved in Alzheimer's disease. Proc Natl Acad Sci USA. 2004;101:10804‐10809. doi: 10.1073/PNAS.0400348101 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 325. Serenó L, Coma M, Rodríguez M, et al. A novel GSK‐3beta inhibitor reduces Alzheimer's pathology and rescues neuronal loss in vivo. Neurobiol Dis. 2009;35:359‐367. doi: 10.1016/J.NBD.2009.05.025 [DOI] [PubMed] [Google Scholar]
- 326. Hampel H, Ewers M, Bürger K, et al. Lithium trial in Alzheimer's disease: a randomized, single‐blind, placebo‐controlled, multicenter 10‐week study. J Clin Psychiatry. 2009;70:10490. doi: 10.4088/JCP.08M04606 [DOI] [PubMed] [Google Scholar]
- 327. Tolosa E, Litvan I, Höglinger GU, et al. A phase 2 trial of the GSK‐3 inhibitor tideglusib in progressive supranuclear palsy. Mov Disord. 2014;29:470‐478. doi: 10.1002/MDS.25824 [DOI] [PubMed] [Google Scholar]
- 328. Leclair‐Visonneau L, Rouaud T, Debilly B, et al. Randomized placebo‐controlled trial of sodium valproate in progressive supranuclear palsy. Clin Neurol Neurosurg. 2016;146:35‐39. doi: 10.1016/J.CLINEURO.2016.04.021 [DOI] [PubMed] [Google Scholar]
- 329. VandeVrede L, Dale ML, Fields S, et al. Open‐label phase 1 futility studies of salsalate and young plasma in progressive supranuclear palsy. Mov Disord Clin Pract. 2020;7:440‐447. doi: 10.1002/MDC3.12940 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 330. Lagraoui M, Sukumar G, Latoche JR, Maynard SK, Dalgard CL, Schaefer BC. Salsalate treatment following traumatic brain injury reduces inflammation and promotes a neuroprotective and neurogenic transcriptional response with concomitant functional recovery. Brain Behav Immun. 2017;61:96‐109. doi: 10.1016/J.BBI.2016.12.005 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 331. Min SW, Chen X, Tracy TE, et al. Critical role of acetylation in tau‐mediated neurodegeneration and cognitive deficits. Nat Med. 2015;21:1154‐1162. doi: 10.1038/nm.3951 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 332. Prevail Therapeutics Announces FDA Orphan Drug Designation Granted to PR006 for the Treatment of Patients with Frontotemporal Dementia with a GRN Mutation | BioSpace n.d. https://www.biospace.com/article/releases/prevail‐therapeutics‐announces‐fda‐orphan‐drug‐designation‐granted‐to‐pr006‐for‐the‐treatment‐of‐patients‐with‐frontotemporal‐dementia‐with‐a‐grn‐mutation/ (accessed December 6, 2023)
- 333. ASN120290 receives Orphan Drug Designation from the FDA | Asceneuron n.d. https://www.asceneuron.com/news/asn120290‐receives‐orphan‐drug‐designation‐fda (accessed December 6, 2023)
- 334. Alectos Therapeutics Announces FDA Orphan Drug Designation n.d. https://www.globenewswire.com/en/news‐release/2016/04/20/830758/0/en/Alectos‐Therapeutics‐Announces‐FDA‐Orphan‐Drug‐Designation‐for‐MK‐8719‐An‐Investigational‐Small‐molecule‐OGA‐Inhibitor‐for‐Treatment‐of‐Progressive‐Supranuclear‐Palsy.html (accessed December 6, 2023)
- 335. Espíndola SL, Damianich A, Alvarez RJ, et al. Modulation of tau isoforms imbalance precludes tau pathology and cognitive decline in a mouse model of tauopathy. Cell Rep. 2018;23:709‐715. doi: 10.1016/J.CELREP.2018.03.079 [DOI] [PubMed] [Google Scholar]
- 336. Muñiz JA, Facal CL, Urrutia L, et al. SMaRT modulation of tau isoforms rescues cognitive and motor impairments in a preclinical model of tauopathy. Front Bioeng Biotechnol. 2022;10:951384. doi: 10.3389/FBIOE.2022.951384/BIBTEX [DOI] [PMC free article] [PubMed] [Google Scholar]
- 337. Avale ME, Rodríguez‐Martín T, Gallo JM. Trans‐splicing correction of tau isoform imbalance in a mouse model of tau mis‐splicing. Hum Mol Genet. 2013;22:2603‐2611. doi: 10.1093/HMG/DDT108 [DOI] [PMC free article] [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.
Supplementary Materials
Supporting information
