ArticlePDF AvailableLiterature Review

Tau-mediated neurodegeneration in Alzheimer's disease and related disorders

Authors:

Abstract

Advances in our understanding of the mechanisms of tau-mediated neurodegeneration in Alzheimer's disease (AD) and related tauopathies, which are characterized by prominent CNS accumulations of fibrillar tau inclusions, are rapidly moving this previously underexplored disease pathway to centre stage for disease-modifying drug discovery efforts. However, controversies abound concerning whether or not the deleterious effects of tau pathologies result from toxic gains-of-function by pathological tau or from critical losses of normal tau function in the disease state. This Review summarizes the most recent advances in our knowledge of the mechanisms of tau-mediated neurodegeneration to forge an integrated concept of those tau-linked disease processes that drive the onset and progression of AD and related tauopathies.
© 2007 Nature Publishing Group
A growing body of evidence suggests that the genera-
tion of proteinaceous aggregates is a common patho-
logical process in numerous neurodegenerative diseases.
Indeed, not only is the defining characteristic of several
neurodegenerative diseases the accumulation of pro-
teinacious fibrillary substances (such as senile plaques
(SPs) made of β-amyloid (Aβ), or neurofibrillary tan-
gles (NFTs) made of tau), but significant circumstantial
evidence also clearly implicates these aggregates in the
onset and progression of most aging-related neuro-
degenerative disorders that manifest clinically with
progressive cognitive and/or motor impairments. In
the case of neurodegenerative tauopathies a group of
disorders that includes Alzheimer’s disease (AD) and the
frontotemporal dementias (FTDs) NFTs consisting
of aggregated straight or paired helical filaments (SFs
and PHFs, respectively), twisted ribbons or other con-
formations
1
of aberrantly phosphorylated forms of the
microtubule-associated protein (MAP) tau are the diag
-
nostic hallmark lesions in the CNS. Although the precise
role of these and other specific diagnostic lesions in the
different stages of neurodegenerative disease pathology
is not yet fully understood, it is increasingly evident
that tau-mediated neurodegeneration may result from
the combination of toxic gains-of-function acquired
by the aggregates or their precursors and the detrimental
effects that arise from the loss of the normal function(s)
of tau in the disease state. Elucidating the exact roles of
the different aggregates and their precursors in neuro-
degeneration is a challenging endeavor, but one that is
likely to remain the focus of future research efforts to
discover the mechanisms of disease pathology, as well as
to develop better diagnostics and therapeutics.
Thus far, several lines of investigation have suggested
different, and at times contradictory, cause-and-effect
relationships between various pathological species
of disease proteins and the aggregates that they form.
These apparent contradictions may reflect the limita-
tions inherent in each of the in vitro and in vivo model
systems that are used to study specific disorders, includ-
ing neurodegenerative tauopathies. For example, the
realization that the neurotoxic species that contribute
to disease onset and progression may be ‘hiddenin one
or more of the pre-aggregated/pre-fibrillar forms of the
misfolded protein clearly complicates both experimental
design and the unambiguous interpretation of results.
Moreover, added complexity may come from the fact
that, aside from its well-established role in promoting
the stabilization of microtubules (MTs), tau may have
additional functions as a result of its interactions with
other structures and enzymes
1
(for example, with the
plasma membrane
2,3
, the actin cytoskeleton
4
and with src
tyrosine kinases such as FYN
5
(see below)). Such poorly-
defined interactions and functions of tau contribute to
the difficulty of understanding how pathologically altered
tau mediates neurodegeneration, and more studies
*Center for
Neurodegenerative Disease
Research, Department of
Pathology and Laboratory
Medicine, University of
Pennsylvania, 3600 Spruce
Street, Philadelphia,
Pennsylvania 19104-4283,
USA.
Institute on Aging,
University of Pennsylvania,
3615 Chestnut Street,
Philadelphia, Pennsylvania
19104-2676, USA.
§
Department of Chemistry,
University of Pennsylvania,
231 South 34th Street,
Philadelphia, Pennsylvania
19104-6323, USA.
Correspondence to J.Q.T.
e-mail: trojanow@mail.med.
upenn.edu
doi:10.1038/nrn2194
Published online
8 August 2007
Senile plaque
A site of Aβ accumulation and
dystrophic neurites in the
brains of mouse models and
patients with Alzheimer’s
disease.
Tau-mediated neurodegeneration
in Alzheimers disease and related
disorders
Carlo Ballatore*
§
, Virginia M.-Y. Lee*
and John Q. Trojanowski*
Abstract | Advances in our understanding of the mechanisms of tau-mediated
neurodegeneration in Alzheimer’s disease (AD) and related tauopathies, which are
characterized by prominent CNS accumulations of fibrillar tau inclusions, are rapidly moving
this previously underexplored disease pathway to centre stage for disease-modifying drug
discovery efforts. However, controversies abound concerning whether or not the deleterious
effects of tau pathologies result from toxic gains-of-function by pathological tau or from
critical losses of normal tau function in the disease state. This Review summarizes the most
recent advances in our knowledge of the mechanisms of tau-mediated neurodegeneration
to forge an integrated concept of those tau-linked disease processes that drive the onset and
progression of AD and related tauopathies.
NATURE REVIEWS
|
NEUROSCIENCE VOLUME 8
|
SEPTEMBER 2007
|
663
REVIEWS
© 2007 Nature Publishing Group
Nature Reviews | Neuroscience
Mechanisms possibly involved in
tau hyperphosphorylation
Direct events
Upregulation or abberant
activation of tau kinases
Downregulation of phosphatases
Mutations
Covalent modifications of tau
Others?
Indirect events
Aβ-mediated toxicity
Oxidative stress
Inflammation
Others?
Neurotoxicity
Compromised axonal
transport
Loss of function
Toxic gains-of-function
NFTs made of hyperphosphorylated
tau sequester normal tau
Hyperphosphorylated tau
Detachment from MTs;
loss of MT-stabilizing function
Changes in
MT dynamics
NFTs become physical obstacles to the
transport of vesicles and other cargoes
Alternative splicing
The process by which introns
are excised from RNA after
transcription and the cut ends
of the RNA are rejoined to form
a continuous message.
Alternative splicing allows the
production of different
messages from the same DNA
molecule.
are needed in order to elucidate these mechanisms. In
addition, because disease onset and progression are
dynamic processes that take place over time (often over
several years), it is conceivable that processes such as
the aggregation of altered tau may produce a range of
different effects at various stages of the disease.
Given the complexities of this research, it is timely to
critically analyse the progress made towards a mechanis-
tic understanding of tau-mediated neurodegeneration,
and to discuss the therapeutic strategies that target the
most severe toxic consequences of tau pathologies. To
that end, our goal is to summarize the current under-
standing of normal tau functions and the pathogenesis
of tau aggregates in AD and related neurodegenerative
tauopathies, as well as their significance to the onset and
progression of these disorders (FIG. 1). We also provide an
overview of the animal models of tau-mediated neuro
-
degeneration (BOX 1; TABLE 1), and discuss tau-directed
drug-discovery efforts (
BOX 2; TABLE 2).
Physiological functions of tau
The primary function of the MAP tau, which is par-
ticularly abundant in the axons of neurons, is to stabi-
lize MTs. As summarized in FIG. 2, there are six major
isoforms of tau expressed in the adult human brain, all
of which are derived from a single gene by alternative
splicing. From a structural stand-point, tau is character-
ized by the presence of a MT-binding domain, which
is composed of repeats of a highly conserved tubulin-
binding motif
6
and which comprises the carboxy-
terminal (C-terminal) half of the protein, followed by a
basic proline-rich region and an acidic amino-terminal
(N-terminal) region, which is normally referred to as the
projection domain. The six tau isoforms differ from each
other in the number of tubulin-binding repeats (either
three or four, hence the isoforms are normally referred
to as 3R and 4R tau isoforms, respectively) and in the
presence or absence of either one or two 29 amino-acid-
long inserts at the N-terminal portion of the protein,
which is not instrumental for MT-binding
7
. Although
the six isoforms appear to be broadly functionally
similar, each is likely to have precise, and to some extent
distinctive, physiological roles. The various isoforms
appear to be differentially expressed during develop-
ment, however, the 3R and 4R tau isoforms are expressed
in a one-to-one ratio in most regions of the adult
brain, and deviations from this ratio are characteristic
of neurodegenerative FTD tauopathies
8
.
Several lines of investigation substantiate a model
whereby the tubulin-binding repeats bind to specific
pockets in β-tubulin at the inner surface of the MTs,
whereas the positively charged proline-rich regions are
Figure 1 | Direct and indirect pathological events that can contribute to tau-mediated neurodegeneration.
Pathological events that can contribute to tau-hyperphosphorylation and detachment from microtubules are shown in the
box on the left. The middle box shows the mechanisms that underlie the loss of normal function and toxic gain-of-function
of tau, which ultimately result in impaired axonal transport and lead to synaptic dysfunction and neurodegeneration (right
hand box). Aβ, amyloid-β; MT, microtubule; NFT, neurofibrillary tangle.
Box 1 | Animal models of tau pathology
The use of invertebrates and rodents in reproductions of
human neurodegenerative diseases that affect higher
cognitive functions over an extended period of time
(decades), although faced with numerous challenges, has
been undeniably instrumental in gaining insights into the
mechanisms that underlie these diseases
64,75-77
. Indeed,
although none of the available models is capable of
recapitulating all the features of tangle pathology that are
found in the human brain, various transgenic models
(TABLE 1) are capable of reproducing different features of
Alzheimer’s disease (AD) in humans, and triple-transgenic
mice can now reproduce multiple signature lesions such
as plaques and NFTs in the same animal model. As such,
these transgenic animal models comprise the most
effective tool available to date for studying AD and
related tauopathies. Some of the commonly used
transgenic animal models are typically associated with the
development of motor impairments, which can severely
limit the use of these models in behavioural tests.
However, recent lines of research
78
have led to the
development of transgenic tauopathy models that have
no overt signs of motor deficits. These particular models
will certainly be invaluable for evaluating the effects of
candidate drugs on cognitive decline.
REVIEWS
664
|
SEPTEMBER 2007
|
VOLUME 8 www.nature.com/reviews/neuro
© 2007 Nature Publishing Group
tightly bound to the negatively charged MT-surface,
and the negatively charged projection domain
branches away from the MT-surface, possibly owing to
electrostatic repulsion
9,10
. Interestingly, it was found
that the MT-binding domain of tau, and several
MT-stabilizing drugs, including paclitaxel, epothilone
and discodermolide, seem to share, either completely
or in part, the same binding pockets in the β-tubulin
9,11
.
Although the occupation of these pockets by tau, other
MAPs or MT-stabilizing drugs is thought to be sufficient
to maintain the tubulin conformations that promote the
polymerized state, MAPs, unlike MT-stabilizing agents,
may also contribute to MT-stabilization in other ways.
It is believed that the β-tubulin pockets of adjacent
protofilaments may be occupied by the different repeats
of the same MT-binding domain of tau, thereby causing
the crosslinking of three or four dimers
9
. In addition,
interactions of the proline-rich region of tau with the
surface of the MTs are likely to further contribute to
MT stabilization.
Interestingly, although the primary function of the
MT-binding domain of tau is the stabilization of MTs,
various lines of investigation have indicated that it may
also engage with other structures and enzymes, includ-
ing RNA
12
and presenilin 1 (PS1)
13
. Similarly, numerous
possible binding partners have been proposed for both
the proline-rich and the projection domains (the SH3
domains of src-family tyrosine kinases such as FYN
5,
and the plasma membrane
2,3
, respectively). Although
the importance of these specific interactions of tau
with partner structures other than the MTs is not yet
known in the context of tau-mediated neurodegenera
-
tion, collectively these findings support the notion that
tau might be a rather promiscuous binder that is prone
Table 1 | Commonly used transgenic mouse models of tau pathology
Gene Mutation/construct Promoter Tau pathology Refs
Mapt 4R/2N isoform Thy1 Hyperphosphorylated PHFs 82
Axonopathy without formation of neuronal NFTs 83
Axonopathy containing neurofilament- and
tau-immunoreactive spheroids, especially in the
spinal cord
84
Fetal tau (3R/0N isoform) Prion protein promoter NFTs in the brain at 18 months of age 56,85
P301L Prion protein promoter Tangle pathology detectable at 2.5 months of age 60,86
P301L Thy1.2 Tangle pathology detectable at 3 months of age 87
Inducible overexpression
of P301L
Ca
2+
–calmodulin-dependent
kinase II
Tangle pathology detectable at 2.5 months of age 13
Genomic tau Endogenous Tau-immunoreactive axonal swellings 88
G272V Prion protein promoter Oligodendroglial fibrillary lesions 89
P301S Thy1.2 Tau pathology detectable at 5 months of age 90
G272V P301L R406W Thy1 Tau pathology detectable at 1.5 months of age.
No motor impairment observed for up to 12 months
after birth
91
V337M
PDGF-β
Mutant tau induces neuronal degeneration,
associated with the accumulation of RNA and
phosphorylated tau
92
R406W Ca
2+
–calmodulin-dependent
kinase II
Hyperphosphorylated tau inclusions appear
in the forebrain at 18 months of age. No motor
abnormalities for up to 23 months after birth
93
G272V P301S Thy1.2 Hyperphosphorylated tau, tangles and PHFs.
No motor impairment for up to 18 months after birth
78
P301S Prion protein promoter This model recapitulates tauopathy, including early
indications of degeneration, such as synapse loss
and microglia activation
63
Apolipoprotein E (Apoe) ApoE4 Multiple Phosphorylated tau expression in the neocortex,
the hippocampus and the amygdala
94
Cdk5r1 p25 Neuron-specific enolase Phosphorylated tau expression in the cortex,
the amygdala and the thalamus
95
VKαD11HuCk and
VKαD11HuCγ
Anti-NGF IgH/Igk
Cytomegalovirus early region Phosphorylated tau expression in the cortex and
the hippocampus, with associated neuron loss
96
App, Psen1, and Mapt PS1 (M146V), APP (Swe),
tau (P301L)
Thy1.2
Both tau and Aβ pathology
97,98
All transgenic mice overexpressed their transgenic protein. VK
α
D11HuC
k
and VK
α
D11HuC
γ
carried the light and heavy chain genes of the chimeric antibodies
αD11, respectively. Aβ, amyloid-β; App, amyloid-b precursor protein; Ig, immunoglobulin; NFT, neurofibrillary tangle; NGF, nerve growth factor; PDGF-β, platelet-
derived growth factor-β; PHF, paired helical filament; PS1, presenilin 1.
REVIEWS
NATURE REVIEWS
|
NEUROSCIENCE VOLUME 8
|
SEPTEMBER 2007
|
665
© 2007 Nature Publishing Group
to heterogeneous interactions particularly when
disengaged from the MT — which may lead to protein
misfolding and aggregation
14
.
The MT-binding ability of tau is post-translationally
regulated primarily by serine/threonine-directed phos
-
phorylation (FIG. 3), which can effectively modulate the
binding affinity of tau for MTs
15
. This is thought to be
the most prominent mechanism that regulates the affinity
of tau for the MTs
15
, although other post-translational
modifications, such as glycosylation
16–18
, may also have
a direct impact on the dynamic equilibrium of tau on
and off the MTs (see below). Notably, phosphoryla-
tion of tau appears to be developmentally regulated
— it is substantially higher during the development of
the fetal brain. On the other hand, in the adult brain,
neurons are normally characterized by a considerably
lower tau phosphorylation state. Furthermore, in the
course of tau-mediated neurodegeneration, aberrant tau
phosphorylation is always observed (see below).
Aside from phosphorylation and glycosylation, other
post-translational modifications of tau also occur
(see REF. 19 for a recent review), including glycation
20
,
ubiquitylation
21
, sumoylation
22,23
, nitration
24
and
proteolysis
25
. Although it is conceivable that most or all
of these post-translational modifications may take place
at various stages of tau pathology, their significance,
particularly in comparison to the well-established role
of phosphorylation, is yet to be fully characterized.
With its ability to modulate MT-dynamics, tau
contributes directly or indirectly to key structural and reg-
ulatory cellular functions. For example, the action of tau on
the MT network has great importance in maintaining an
appropriate morphology of neurons, the processes of which
typically extend over relatively great distances, making
neurons the most asymmetrical of all cells. Furthermore,
because the MT network is key to the sophisticated trans-
port machinery (FIG. 3) that allows signalling molecules,
trophic factors and other essential cellular constituents,
including organelles (for example mitochondria and
vesicles), to travel along the axons (axonal transport), then
tau clearly has profound effects on axonal transport and,
hence, on the function and viability of neurons and their
highly extended processes
26
. Importantly, under normal
physiological conditions, tau is in a constant dynamic
equilibrium, on and off the MTs. This equilibrium is
thought to be controlled primarily by the phosphorylation
state of tau, which in turn is determined by the actions
of kinases and phosphatases. Indeed, frequent cycles of
binding and detachment of tau from the MTs (corre-
sponding to phosphorylations and dephosphorylations,
respectively) may be needed to allow effective axonal
transport (FIG. 3).
Pathological aggregation of tau
Under pathological conditions, the equilibrium of tau
binding to the MTs is perturbed, resulting in an abnor-
mal increase in the levels of the free (unbound) tau
Table 2 | Therapeutic strategies targeting tau that are currently under investigation
Therapeutic approach Expected effect Current status Refs
Kinase inhibition Prevent the abnormal phosphorylation rate
or state of tau and consequent excessive
disengagement of tau from the MTs
Various stages of preclinical investigation 15,37
Inhibition of tau fibril formation or
dissolution of pre-existing aggregates
Prevent aberrantly phosphorylated
and/or misfolded tau from forming more
organized aggregates
Early stages of drug discovery 67–69,71,99
Activation of chaperone systems Facilitate the clearance of misfolded tau
and/or tau aggregates
Early stages of preclinical investigation 100,101
Stabilization of the MTs Compensate for the loss of taus
MT-stabilizing function, and thereby
sustain axonal transport
Different programmes are at various stages
of development, ranging from preclinical
investigations to Phase II studies
55,57,102
Aβ-directed therapies Prevent Aβ from contributing to
tau-mediated neurodegeneration
Several compounds in Phase I and Phase II
studies
75
Attenuation of inflammation Attenuation of inflammation might
contribute to a slowing of the progression
of the disease.
Different programmes are at various stages of
development, ranging from early preclinical
to Phase III studies
63,103–105
Aβ, amyloid-β; MT, microtubule.
Box 2 | The development of therapeutics
The need for therapies that are capable of modifying both amyloid-β (Aβ)- and tau-
mediated neurodegeneration cannot be overemphasized. Although a relatively large
number of Aβ-directed therapeutic approaches have been proposed, and although
many of these are at various stages of clinical investigation
79
, tau-directed therapies
have been lagging behind. Nonetheless, in light of the fact that two of the most
extensively studied cancer targets, namely kinases and MTs, are clearly involved in
tau-mediated neurodegeneration, the search for therapeutics that are capable of
modifying tau-pathology will be able to take advantage of past experience and,
importantly, of the relatively large number of biologically active compounds that have
already been developed. In addition, thanks to the recent development of various
in vitro tau fibrillization assays, high-throughput screening (HTS) of compound libraries
has been possible, and this has led to the identification of structures that are capable
of functioning as imaging ligands to detect aggregates in living patients, inhibiting
tau fibril formation and/or dissolving preformed fibrils
66–69,80
(TABLE 2). Although these
HTS efforts are clearly part of rather early-stage drug-discovery programmes, the hits
discovered thus far have promise as novel molecular tools to further investigate the
pathophysiology of tau. However, in order to conduct thorough evaluations of the
existing therapeutic agents, as well as of novel candidate compounds, in the context
of neurodegenerative diseases, key pharmacokinetic issues such as drug uptake in
the brain, which is notoriously hampered by the presence of the highly discriminating
blood–brain barrier
81
, remain to be resolved.
REVIEWS
666
|
SEPTEMBER 2007
|
VOLUME 8 www.nature.com/reviews/neuro
© 2007 Nature Publishing Group
Nature Reviews | Neuroscience
R1 R2 R3 R4
C terminus
4411
410
381
383
352
412
4R/2N
3R/2N
4R/1N
3R/1N
4R/0N
3R/0N
N terminus
N terminus
N terminus
N terminus
N terminus
N terminus
C terminus
C terminus
C terminus
C terminus
C terminus
1
1
1
1
1
fraction. It is likely that the resultant higher cytosolic
concentrations of tau increase the chances of patho-
genic conformational changes that in turn lead to the
aggregation and fibrillization of tau
14
(FIG. 4). Important
progress has been made in recent years in understand-
ing tau misfolding and fibril formation
14,27,28
. The path
from normal tau bound to the MTs to large aggregate
structures such as NFTs is thought to be a multi-step
phenomenon which begins with the detachment of tau
from the MTs. The key steps of tau fibrillization are
highlighted in FIG. 4. On abnormal disengagement of
tau from the MTs, which can be triggered by numer-
ous causes (including increased rate of phosphoryla-
tion and/or decreased rate of dephosphorylation),
the cytosolic concentration of unbound tau would
rise. This is likely to be a crucial step, one which may
render tau considerably more likely to undergo mis-
folding and may, as a result, make it more prone to
aggregation. Next, small nonfibrillary tau deposits
(normally referred to as ‘pretangles’) are formed, and
these, unlike NFTs, cannot be detected by β-sheet-
specific dyes
29–31
. This indicates that pretangles do not
contain β-sheets, and that a structural rearrangement
involving the formation of the characteristic pleated
β-sheet must occur during the transition from pre-
tangles to PHFs. Finally, PHFs further self-assemble
to form NFTs.
Causes of tau abnormalities in disease
It is believed that several pathogenic events might con-
tribute, either directly or indirectly, to tau hyperphos-
phorylation, misfolding and aggregation. Perhaps the
most direct cause-and-effect relationship was estab
-
lished by seminal genetic studies that demonstrated
that mutations of the tau gene (MAPT) are causative
of FTD with parkinsonism linked to chromosome-17
(FTDP-17)
32,33
. All cases of FTDP-17 are character-
ized by the presence of filamentous inclusions that
are composed of hyperphosphorylated tau. Such
mutations could lead to the expression of tau mutants
that are: predisposed to assembly into filaments and
therefore able to undergo rapid fibrillization
32,34
; more
readily phosphorylated and/or less prone to dephos-
phorylation
35
; or that show impaired MT binding
properties
8,36
. Furthermore, intronic MAPT muta-
tions, as well as most coding-region mutations in exon
10 (N279K, L284L,
N296, N296N, N296H, S305N
and S305S), may alter the alternative splicing of tau
to perturb the normal one-to-one ratio of the 3R to
4R tau isoforms, with the 4R isoform being overpro
-
duced in most, but not all, instances. Indeed, over 30
different tau gene mutations have been identified in
families with FTDP-17 (recently reviewed in
REF. 32).
Importantly, these genetic studies provide unam-
biguous evidence that tau malfunction is sufficient to
trigger neurodegeneration and dementia even in the
absence of other pathogenic insults.
Additional direct cause-and-effect links have been
established between tau malfunctions and an overall
imbalance in the activity levels or regulation of tau
kinases and phosphatases
15,37
. Under physiological
conditions, single tau molecules are typically phos-
phorylated at a subset of potential phosphate-acceptor
amino-acid residues. During late stage neurodegen
-
eration, the phosphorylation state of a single tau mol-
ecule can reach such high levels that many or most
of these residues are phosphorylated and, at the same
time, a higher proportion of tau molecules are in this
hyperphosphorylated state. Although several kinases
have been found to be capable of phosphorylating
tau in vitro, it is not yet clear whether all of them
participate in tau phosphorylation under physiologi-
cal or pathological conditions in vivo
1
. Nonetheless,
glycogen synthase kinase 3 (GSK3), cyclin-dependent
kinase 5 (CDK5) and the microtubule-affinity-regulating
kinase (MARK) have received particular attention as
potential targets for disease-modifying therapies using
inhibitory compounds
15
. For example, inhibition of
GSK3 by lithium not only reduced tau phosphoryla-
tion in vivo, but also lowered the level of aggregated
tau, compared with controls
38
. Other studies on the
effects of lithium in transgenic AD mice also suggested
a concomitant reduction in Aβ production, possibly
resulting from lithium-mediated inhibition of
GSK3α,
which is required for maximal processing of the pre-
cursor of Aβ, amyloid precursor protein (APP)
39
.
Similarly, a number of phosphatases
40
, including pro-
tein phosphatase (PP)1, PP2A, PP2B and PP2C, have
been identified that could potentially drive the reverse,
Figure 2 | The domain structure of the tau isoforms that are expressed in the adult
human brain.
The isoforms can differ from each other in the number of tubulin-binding
domains (three or four repeats located in the C-terminal half of the protein, shown in
red), and are referred to as 3R or 4R tau isoforms, respectively. They can also differ in the
presence or absence of either one or two 29-amino-acid-long, highly acidic inserts
(shown in yellow) at the N-terminal portion of the protein (the projection domain).
Between the projection domain and the microtubule-binding domain lies a basic
proline-rich region.
REVIEWS
NATURE REVIEWS
|
NEUROSCIENCE VOLUME 8
|
SEPTEMBER 2007
|
667
© 2007 Nature Publishing Group
Nature Reviews | Neuroscience
Cargo moving
along the axon
Cargo moving
along the axon
Motor protein
Vesicle
Tau attached
to the MT
MT
Kinase-mediated phosphorylation
detaches tau from the MT
Dephosphorylation reaction by
phosphatases restores the
MT-binding ability of tau
Dephosphorylation
Phosphorylation
Phosphorylated tau
disengaged from the MT
Oxidative stress
A disturbance in the pro-
oxidant–antioxidant balance in
favour of the pro-oxidant,
leading to potential cellular
damage. Indicators of oxidative
stress include damaged DNA
bases, protein oxidation and
lipid peroxidation products.
Dystrophic neurites
The processes (axons and
dendrites) of neurons that are
damaged or degenerating in
AD.
dephosphorylation of tau; however, the exact role of
these phosphatases under physiological and pathologi-
cal conditions is not completely clear.
The overall effect of the increased rate and/or state
of phosphorylation appears to be the abnormal dis-
engagement of tau from the MTs. Furthermore, it is
likely that various other pathological events, includ-
ing Aβ-mediated toxicity, as well as oxidative stress and
inflammation, may be able to trigger or contribute
(independently or in combination) to an abnormal
detachment of tau from the MTs
41–45
. For example, it has
been suggested that oxidative stress could be responsi-
ble for detrimental covalent modifications of tau, which
include the formation of intermolecular disulphide
bridges
46
and tyrosine nitration
24
. Such modifications
are likely to cause misfolding, hyperphosphorylation and
aggregation, and thereby contribute to abnormal disen-
gagement of tau from MTs, as well as to the formation
of aggregates. However, despite the clear involvement of
these pathological processes in tau-mediated neuro-
degeneration, their relative positioning in the cascade
of events that leads to neuronal loss remains unclear. For
example, although oxidative stress is often regarded as an
upstream event relative to tau pathology, recent studies
have revealed that pathological tau may interfere with
mitochondrial function and induce oxidative stress
47
.
This raises the interesting possibility that although oxida-
tive stress is likely to be a relatively early event that could
lead to tau malfunction, it is equally possible that tau
malfunction, once initiated, may further exacerbate the
effects of this, and possibly other, upstream events.
Connections between Aβ-mediated toxicity and tau
pathology have repeatedly been proposed
48,49
; however,
the mechanism or mechanisms that link SPs and NFTs
have not yet been fully established, and this remains one
of the most challenging conundrums of AD research.
Nonetheless, new lines of investigation support the
notion that tau malfunction, in addition to being inde-
pendently capable of producing neurodegeneration
even in the absence of Aβ deposits or other pathological
events
32,33
, could be a key mediator of neurodegen-
eration in response to other upstream events, including
Aβ-induced toxicity
44
. An interesting and unexpected
development of the proposed pathological role of tau as a
common mediator of neurodegeneration is the hypoth-
esis that suppression of tau may potentially be beneficial.
In accordance with this hypothesis, a recent study
50
has
shown that reduction or elimination of endogenous tau,
in a mouse model of AD-like Aβ amyloidosis which
expresses human amyloid precursor protein (hAPP)
with a familial AD mutation that increases Aβ produc-
tion, is beneficial against Aβ-induced deficits. These
results appear to substantiate previous cell-based studies
which showed that cultured hippocampal neurons from
tau-knockout mice treated with fibrillar A
β were not
susceptible to Aβ-induced toxicity
44
. However, although
the most valid model for comparisons with tau suppres-
sion would be a tau-knockdown mouse, it is notable that
tau-knockout mice show behavioural impairments and
structural abnormalities with advancing age
51
, suggest-
ing that long-term suppression of tau as a therapy for
tauopathies might be fraught with complications.
Tau-mediated neurodegeneration
As described above, in AD and related neurodegen-
erative disorders that are collectively referred to as
tauopathies,
52,53
tau no longer binds to the MTs; instead
it becomes sequestered into NFTs in neurons, and into
glial tangles in astrocytes or oligodendroglia. In AD at
least, the largest burden of tau pathology (~95% of total
tau by morphometic analyses) is found in neuronal proc-
esses known as neuropil threads or dystrophic neurites
54
.
In broad terms, the pathological consequences of these
events could result from a loss of normal tau func-
tion combined with gains of pathological functions of
hyperphosphorylated tau, the filaments formed thereof,
and the aggregation of these filaments to form glial and
neuronal tangles in dystrophic neurites.
The loss of tau’s normal MT-stabilizing function
would invariably lead to a pathological disturbance
Figure 3 | The dynamic equilibrium of tau microtubule (MT) binding. A schematic representation of the normal
dynamic equilibrium of tau, on and off the MTs, which is primarily determined by the phosphorylation state of tau.
Although the presence of tau on the MTs presents a physical obstacle for vesicles and other cargoes that are moving along
the axon, MT-bound tau is essential to MT integrity. Thus, relatively frequent cycles of tau–MT binding (promoted by
dephosphorylation of tau) and detachment of tau from the MT (promoted by phosphorylation of tau) are needed in order
to maintain effective axonal transport.
REVIEWS
668
|
SEPTEMBER 2007
|
VOLUME 8 www.nature.com/reviews/neuro
© 2007 Nature Publishing Group
Nature Reviews | Neuroscience
Possible causes of tau aggregation
Tau gene
mutations
Covalent
modifications
of tau
Others?
Phosphorylation
(by kinases)
Dephosphorylation
(by phosphatases)
Misfolded tau
Pretangles
β-sheet
containing structures
(PHFs)
Filamentous inclusions
(NFTs)
a
b
c
d
e
Detachment of tau from the MTs. Increased unbound tau.
Microglia
A non-neuronal cell type that is
present in the spinal cord and
the brain (it is the resident CNS
macrophage) and is
characterized by its ramified
morphology.
in the normal structural and regulatory functions of
the cytoskeleton, which would compromise axonal
transport and thus contribute to synaptic dysfunction
and neurodegeneration
26,55
. Indeed, the importance of
the loss of the MT-stabilizing function of tau in neuro-
degeneration was recently validated by proof-of-concept
studies carried out in vivo, which demonstrated that the
MT-stabilizing drug paclitaxel can ameliorate the neuro-
degenerative phenotype of transgenic mouse models
of AD-like tau amyloid pathologies
56,57
. However, the
discovery that the total level of NFTs correlates with
the degree of cognitive impairment
58,59
provided the ini-
tial circumstantial evidence to suggest that toxic gains-
of-function by NFTs might play an important part in the
progression of the disease. Indeed, pioneering studies
that used immunohistochemical techniques to deter-
mine the level of both NFTs and SPs in different brain
regions of AD patients, as well as non-demented elderly
individuals, demonstrated that the number of NFTs, but
not the numbers of SPs, correlates with the degree of
cognitive impairment
58,59
.
It is possible that the toxic effects of NFTs may
partly arise from the relatively large size of the fibril-
lary material that accumulates inside the neurons,
as this material may pose a direct physical disrup-
tion to cellular functions such as axonal transport.
Furthermore, NFTs may also contribute to the disease
progression by effectively sequestering more tau and
other proteins, and thereby reinforcing and amplifying
the loss of normal tau function.
However, the notion that NFTs could have a promi-
nent role in the progression of the disease was recently
challenged by reports that suppression of transgenic tau
in a neurodegenerative tauopathy mouse model pro-
duced improvements in memory function, even though
NFTs continued to accumulate
60,61
. However, it should
be noted that in the model used, the degree of tau sup-
pression is relative to the fully activated state of tau: this
means that a 2.5-fold overexpression of tau (compared
with endogenous tau) would still be present. Another
significant observation was that, in a mouse model of
AD-like Aβ pathology, axonal defects that consisted
of swellings that contained accumulated abnormal
amounts of tau, other proteins and vesicles were found
to precede the appearance of Aβ deposits, including
SPs, by more than 1 year
62
. Furthermore, restoration of
impaired axonal transport in a tauopathy mouse model
rescued the disease phenotype
57
. In addition, studies in
a transgenic (P301S) tauopathy mouse model revealed
that synapse loss and microglial activation precede the
appearance of NFTs, presumably due to the impaired
transport that results from tau hyperphosphorylation
63
.
Collectively, these studies substantiate the notion that
axonal transport defects, synapse loss and neuroinflam-
mation may be among the earliest signs of neurodegen-
eration that results from tau hyperphosphorylation,
whereas fibrillary tau tangles may be late-stage manifes
-
tations that could contribute to the disease progression
by physically interfering with normal cellular functions.
At the same time, the tangles may sequester larger
quantities of other functionally significant proteins,
and thereby exacerbate and amplify upstream causes. It
should be noted that the possible relative contribution
to neurodegeneration from the toxic gains-of-function
versus the loss of normal function may be experimen-
tally difficult to discern, as the toxic gain may imply, at
least in part, an amplification of the loss of function. In
addition, a precise correlation between the size of the
NFTs and these toxic gains (that is, the putative criti-
cal mass that is required for an insoluble intracellular
deposit to become a physical obstacle) is not yet known.
Figure 4 | Pathological aggregation of tau. A schematic
representation of the different stages of the formation of
pathological tau aggregates. a | Abnormal disengagement
of tau from the MTs and a concomitant increase in the
cytosolic concentration of tau are likely to be the key
events that lead to tau-mediated neurodegeneration.
Direct causes of abnormal disengagement of tau from the
MTs include an imbalance of tau kinases and/or
phosphatases, mutations of the tau gene, covalent
modification of tau causing and/or promoting misfolding,
and possibly other causes such as other post-translational
modifications. b | Once tau is unbound from the MT it
becomes more likely to misfold. This is thought to be a
stochastic phenomenon that is more likely at higher
cytosolic tau concentrations. c | Early deposits of tau,
called ‘pretangles’, are not stained by congo red or
thioflavine-T, indicating that these intermediate forms
of aggregated tau do not exhibit the pleated β-sheet
structure typically found in amyloid aggregates.
d
| A structural transition leads to this more organized
aggregate and the eventual development of neurofibrillary
tangles (e). Such transitions may be facilitated by
heterogeneous interactions with membranous
structures
14,29
. MT, microtubule; NFT, neurofibrillary tangle;
PHF, paired helical filament.
REVIEWS
NATURE REVIEWS
|
NEUROSCIENCE VOLUME 8
|
SEPTEMBER 2007
|
669
© 2007 Nature Publishing Group
Furthermore, with the existence of pre-fibrillary tau spe-
cies
30,31
, toxic gains-of-function by abnormal tau could be
ascribed to one or more of these ill-defined intermediate
species. Thus, although loss of tau function and the toxic
gain-of-function by PHFs or other abnormal species of
tau, in addition to the toxic properties acquired by NFTs
as they enlarge, may contribute to neurodegeneration to
different extents, it is highly plausible that both types of
mechanism contribute to the onset and progression
of AD and other tauopathies, especially at different
stages of the pathology.
Conclusions and future directions
Despite significant recent advances in our under-
standing of tau-mediated neurodegeneration, which
substantiate the notion that tau may act as a common
mediator of neurodegeneration for various upstream
pathological events, a detailed picture of causes and
effects has not yet emerged. Thus, although it is increas-
ingly clear that the disengagement of tau from MTs is
likely to comprise a cardinal step that sets the stage for
tau-mediated neurodegeneration, the links between
this and other upstream events such as Aβ-mediated
toxicity and oxidative stress remain less clear. Likewise,
although tau-mediated neurodegeneration probably
results from the combination of losses of function and
toxic gains-of-function, the specific roles played by the
various forms of misfolded and aggregated tau are not
fully understood. It should be noted, however, that the
onset and progression of the disease may not necessarily
be best represented by an unequivocal linear sequence
of causes and effects, where one single ‘root’ cause is
responsible for the entire cascade of events. With the
recognition that amplification mechanisms exist that
could effectively short-circuit cause and effect and
thereby exacerbate each-others’ detrimental effects,
it is plausible that disease progression may lie in the
ability to set in motions such self-sustaining cycles.
Thus, continuing efforts should be made to further
characterize the precise mechanism(s) by which differ-
ent pathological events may influence and amplify each
other. With a wide range of animal models that partially
recapitulate the key phenotypic features of AD and
related tauopathies already developed
64
(BOX 1; TABLE 1),
further insight into the mechanistic roles of the different
species of tau aggregates in neurodegenerative disease
relies on the discovery of a wider set of molecular tools.
These tools must be capable of selectively modulating
and/or imaging
65
putative pathogenic and pathologi-
cal steps. To this end, functional genomics and drug
discovery efforts aimed at some of these targets will
be instrumental (BOX 2; TABLE 2). For example, agents
that are capable of slowing down, blocking or reversing
protein tau aggregation in vitro
66–74
have been identified;
however, issues such as lack of selectivity and/or toxicity
have limited their use in vivo. In addition, several agents
that could compensate for the loss of tau function, such
as paclitaxel and other MT-stabilizing agents
55
that have
been part of the medical armamentarium for several
years, are hampered by limited CNS uptake and tox-
icities. Drug discovery and lead optimization efforts
that could succeed in making these and other agents
available for in vivo testing will greatly facilitate further
understanding of the tau pathway, as well as of the
pathogenic significance of specific events, whose exact
relation with tau malfunction is less defined.
1. Buee, L., Bussiere, T., Buee-Scherrer, V., Delacourte, A.
& Hof, P. R. Tau protein isoforms, phosphorylation and
role in neurodegenerative disorders. Brain Res. Rev.
33, 95–130 (2000).
2.
Brandt, R., Leger, J. & Lee, G. Interaction of tau with
the neural plasma membrane mediated by tau’s
amino-terminal projection domain. J. Cell Biol. 131,
1327–1340 (1995).
3.
Maas, T., Eidenmuller, J. & Brandt, R. Interaction of
tau with the neural membrane cortex is regulated by
phosphorylation at sites that are modified in paired
helical filaments. J. Biol. Chem. 275, 15733–15740
(2000).
4.
Fulga, T. A. et al. Abnormal bundling and
accumulation of F-actin mediates tau-induced
neuronal degeneration in vivo. Nature Cell Biol. 9,
139–148 (2007).
5.
Lee, G. Tau and src family tyrosine kinases.
Biochim. Biophys. Acta 1739, 323–330 (2005).
6.
Lee, G., Neve, R. L. & Kosik, K. S. The microtubule
binding domain of tau protein. Neuron 2, 1615–1624
(1989).
7.
Binder, L. I., Frankfurter, A. & Rebhun, L. I.
The distribution of tau in the mammalian central
nervous system. J. Cell Biol. 101, 1371–1378 (1985).
8.
Hong, M. et al. Mutation-specific functional
impairments in distinct tau isoforms of hereditary
FTDP-17. Science 282, 1914–1917 (1998).
9.
Amos, L. A. Microtubule structure and its stabilisation.
Org. Biomol. Chem. 2, 2153–2160 (2004).
10.
Kar, S., Fan, J., Smith, M. J., Goedert, M. & Amos, L. A.
Repeat motifs of tau bind to the insides of
microtubules in the absence of taxol. EMBO J. 22,
70–77 (2003).
11.
Kar, S., Florence, G. J., Paterson, I. & Amos, L. A.
Discodermolide interferes with the binding of tau
protein to microtubules. FEBS Lett. 539, 34–36
(2003).
12.
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. 451, 39–44 (1999).
13.
Takashima, A. et al. Presenilin 1 associates with
glycogen synthase kinase-3β and its substrate tau.
Proc. Natl Acad. Sci. USA 95, 9637–9641 (1998).
14.
Kuret, J. et al. Evaluating triggers and enhancers of tau
fibrillization. Microsc. Res. Tech. 67, 141–155 (2005).
This review provides a model to rationalize the
multistep pathway to tau fibril formation, as well as
experimental methods for tau fibrillization assays.
15.
Mazanetz, M. P. & Fischer, P. M. Untangling tau
hyperphosphorylation in drug design for
neurodegenerative diseases. Nature Rev. Drug Discov.
6, 464–479 (2007).
An up‑to‑date account of the role of specific kinases
in tau‑mediated neurodegeneration and their
significance as targets for therapeutic intervention.
16.
Arnold, C. S. et al. The microtubule-associated protein
tau is extensively modified with O-linked
N-acetylglucosamine. J. Biol. Chem. 271,
28741–28744 (1996).
17.
Li, X., Lu, F., Wang, J.-Z. & Gong, C.-X. Concurrent
alterations of O-GlcNAcylation and phosphorylation
of tau in mouse brains during fasting. Eur. J. Neurosci.
23, 2078–2086 (2006).
18.
Liu, F., Iqbal, K., Grundke-Iqbal, I., Hart, G. W. &
Gong, C.-X. O-GlcNAcylation regulates phosphorylation
of tau: a mechanism involved in Alzheimer’s disease.
Proc. Natl Acad. Sci. USA 101, 10804–10809 (2004).
19.
Gong, C. X., Liu, F., Grundke-Iqbal, I. & Iqbal, K.
Post-translational modifications of tau protein in
Alzheimer’s disease. J. Neural Transm. 112, 813–838
(2005).
20.
Münch, G., Deuther-Conrad, W. & Gasic-Milenkovic, J.
Glycoxidative stress creates a vicious cycle of
neurodegeneration in Alzheimer’s disease – a target
for neuroprotective treatment strategies? J. Neural
Transm. 62 (Suppl.), 303–307 (2002).
21.
Cripps, D. et al. Alzheimer disease-specific
conformation of hyperphosphorylated paired helical
filament-tau is polyubiquitinated through Lys-48,
Lys-11, and Lys-6 ubiquitin conjugation. J. Biol. Chem.
281, 10825–10838 (2006).
22.
Dorval, V. & Fraser, P. E. Small ubiquitin-like modifier
(SUMO) modification of natively unfolded proteins tau
and α-synuclein. J. Biol. Chem. 281, 9919–9924
(2006).
23.
Dorval, V. & Fraser, P. E. SUMO on the road to
neurodegeneration. Biochim. Biophys. Acta 1773,
694–706 (2007).
24.
Mailliot, C., Trojanowski, J. Q. & Lee, V. M. Impaired
tau protein function following nitration-induced
oxidative stress in vitro and in vivo. Neurobiol. Aging
23 (Suppl. 1), 415 (2002).
25.
Johnson, G. Tau phosphorylation and proteolysis:
insights and perspectives. J. Alzheimers Dis. 9,
243–250 (2006).
26.
Roy, S., Zhang, B., Lee, V. M.-Y. & Trojanowski, J. Q.
Axonal transport defects: a common theme in
neurodegenerative diseases. Acta Neuropathol. 109,
5–13 (2005).
Reviews the biology of axonal transport and its role
in neurodegenerative disease.
27.
Kuret, J. et al. Pathways of tau fibrillization.
Biochim. Biophys. Acta 1739, 167–178 (2005).
28.
Ross, C. A. & Poirier, M. A. Protein aggregation
and neurodegenerative disease. Nature Med. 10,
S10–S17 (2004).
29.
Galvan, M., David, J. P., Delacourte, A., Luna, J. &
Mena, R. Sequence of neurofibrillary changes in aging
and Alzheimer’s disease: a confocal study with
phospho-tau antibody, AD2. J. Alzheimers Dis. 3,
417–425 (2001).
REVIEWS
670
|
SEPTEMBER 2007
|
VOLUME 8 www.nature.com/reviews/neuro
© 2007 Nature Publishing Group
30. Maeda, S. et al. Granular tau oligomers as
intermediates of tau filaments. Biochemistry 46,
3856–3861 (2007).
31.
Maeda, S. et al. Increased levels of granular tau
oligomers: an early sign of brain aging and
Alzheimer’s disease. Neurosci. Res. 54, 197–201
(2006).
32.
Goedert, M. & Jakes, R. Mutations causing
neurodegenerative tauopathies. Biochim. Biophys.
Acta 1739, 240–250 (2005).
33.
von Bergen, M. et al. Mutations of tau protein in
frontotemporal dementia promote aggregation of
paired helical filaments by enhancing local
β-structure. J. Biol. Chem. 276, 48165–48174
(2001).
34.
Nacharaju, P. et al. Accelerated filament formation
from tau protein with specific FTDP-17 missense
mutations. FEBS Lett. 447, 195–199 (1999).
35.
Alonso Adel, C., Mederlyova, A., Novak, M.,
Grundke-Iqbal, I. & Iqbal, K. Promotion of
hyperphosphorylation by frontotemporal
dementia tau mutations. J. Biol. Chem. 279,
34873–34881 (2004).
36.
Dayanandan, R. et al. Mutations in tau reduce its
microtubule binding properties in intact cells and
affect its phosphorylation. FEBS Lett. 446, 228–232
(1999).
37.
Churcher, I. Tau therapeutic strategies for the
treatment of Alzheimer’s disease. Curr. Top. Med.
Chem. 6, 579–595 (2006).
38.
Noble, W. et al. Inhibition of glycogen synthase
kinase-3 by lithium correlates with reduced tauopathy
and degeneration in vivo. Proc. Natl Acad. Sci. USA
102, 6990–6995 (2005).
Study that validates GSK3β as a target for
tau‑directed therapies.
39.
Phiel, C. J., Wilson, C. A., Lee, V. M. Y. & Klein, P. S.
GSK-3α regulates production of Alzheimer’s
disease amyloid-β peptides. Nature 423, 435–439
(2003).
40.
Tian, Q. & Wang, J. Role of serine/threonine protein
phosphatase in Alzheimer’s disease. Neurosignals 11,
262–269 (2002).
41.
Andersen, J. K. Oxidative stress in
neurodegeneration: cause or consequence?
Nature Med. 5, S18–S25 (2004).
42.
Moreira, P. I. et al. Oxidative stress and
neurodegeneration. Ann. NY Acad. Sci. 1043,
545–552 (2005).
43.
King, M. E. et al. Tau-dependent microtubule
disassembly initiated by prefibrillar β-amyloid. J. Cell
Biol. 175, 541–546 (2006).
44.
Rapoport, M., Dawson, H. N., Binder, L. I., Vitek, M. P.
& Ferreira, A. Tau is essential to β-amyloid-induced
neurotoxicity. Proc. Natl Acad. Sci. USA 99,
6364–6369 (2002).
Study that substantiates the notion that tau may
be a mediator of upstream pathological events,
namely Aβ‑induced neurotoxicity.
45.
Liu, Q. et al. Tau modifiers as therapeutic targets for
Alzheimer’s disease. Biochim. Biophys. Acta 1739,
211–215 (2005).
46.
Schweers, O., Mandelkow, E., Biernat, J. &
Mandelkow, E. Oxidation of cysteine-322 in the
repeat domain of microtubule-associated protein τ
controls the in vitro assembly of paired helical
filaments. Proc. Natl Acad. Sci. USA 92, 8463–8467
(1995).
47.
David, D. C. et al. Proteomic and functional analyses
reveal a mitochondrial dysfunction in P301L tau
transgenic mice. J. Biol. Chem. 280, 23802–23814
(2005).
48.
Blurton-Jones, M. & LaFerla, F. M. Pathways by which
Aβ facilitates tau pathology. Curr. Alzheimer Res. 3,
437–448 (2006).
49.
Oddo, S. et al. Temporal profile of amyloid-β (Aβ)
oligomerization in an in vivo model of Alzheimer
disease: a link between Aβ and tau pathology. J. Biol.
Chem. 281, 1599–1604 (2006).
50.
Roberson, E. D. et al. Reducing endogenous tau
ameliorates amyloid β-induced deficits in an
Alzheimer’s disease mouse model. Science 316,
750–754 (2007).
Study that suggests the role tau might have as a
mediator of neurodegeneration. This study also
suggests that tau reduction may be therapeutically
beneficial.
51.
Ikegami, S., Harada, A. & Hirokawa, N.
Muscle weakness, hyperactivity, and impairment in
fear conditioning in tau-deficient mice. Neurosci. Lett.
279, 129–132 (2000).
52.
Forman, M. S., Trojanowski, J. Q. & Lee, V. M.-Y.
Neurodegenerative diseases: a decade of discoveries
paves the way for therapeutic breakthroughs. Nature
Med. 10, 1055–1063 (2004).
53.
Trojanowski, J. Q. & Mattson, M. P. Overview of
protein aggregation in single, double, and triple
neurodegenerative brain amyloidoses.
Neuromolecular Med. 4, 1–6 (2003).
54.
Mitchell, T. W. et al. Novel method to quantify neuropil
threads in brains from elders with or without cognitive
impairment. J. Histochem. Cytochem. 48,
1627–1638 (2000).
55.
Trojanowski, J. Q., Smith, A. B., Huryn, D. &
Lee, V. M.-Y. Microtubule-stabilizing drugs for therapy
of Alzheimer’s disease and other neurodegenerative
disorders with axonal transport impairments.
Expert Opin. Pharmacother. 6, 683–686 (2005).
56.
Ishihara, T. et al. Age-dependent emergence and
progression of a tauopathy in transgenic mice
overexpressing the shortest human tau isoform.
Neuron 24, 751–762 (1999).
57.
Zhang, B. et al. Microtubule-binding drugs offset tau
sequestration by stabilizing microtubules and
reversing fast axonal transport deficits in a tauopathy
model. Proc. Natl Acad. Sci. USA 102, 227–231
(2005).
58.
Arriagada, P. V., Growdon, J. H., Hedley-Whyte, E. T. &
Hyman, B. T. Neurofibrillary tangles but not senile
plaques parallel duration and severity of Alzheimer’s
disease. Neurology 42, 631–639 (1992).
59.
Arriagada, P. V., Marzloff, K. & Hyman, B. T.
Distribution of Alzheimer-type pathologic changes in
nondemented elderly individuals matches the pattern
in Alzheimer’s disease. Neurology 42, 1681–1688
(1992).
60.
Santacruz, K. et al. Tau suppression in a
neurodegenerative mouse model improves memory
function. Science 309, 476–481 (2005).
Interesting study that shows that suppression of
tau improves memory function even though NFTs
continue to grow.
61.
Trojanowski, J. Q. & Lee, V. M. Pathological tau: a loss
of normal function or a gain in toxicity? Nature
Neurosci. 8, 1136–1137 (2005).
62.
Stokin, G. B. et al. Axonopathy and transport deficits
early in the pathogenesis of Alzheimer’s disease.
Science 307, 1282–1288 (2005).
Together with reference 61, this study shows
that axonal transport defects may be early
pathological events in tau‑mediated
neurodegeneration.
63.
Yoshiyama, Y. et al. Synapse loss and microglial
activation precede tangles in a P301S tauopathy
mouse model. Neuron 53, 337–351 (2007).
Paper demonstrating that microgliosis and
synaptic pathology may be the earliest
manifestation of neurodegenerative tauopathies.
The paper also suggests that abrogation of tau‑
induced microglial activation may be
therapeutically beneficial.
64.
Lee, V. M.-Y., Kenyon, T. K. & Trojanowski, J. Q.
Transgenic animal models of tauopathies. Biochim.
Biophys. Acta 1739, 251–259 (2005).
65.
Shaw, L. M., Korecka, M., Clark, C. M., Lee, V. M. &
Trojanowski, J. Q. Biomarkers of neurodegeneration
for diagnosis and monitoring therapeutics. Nature
Rev. Drug Discov. 6, 295–303 (2007).
66.
Necula, M., Chirita, C. N. & Kuret, J. Cyanine dye
N744 inhibits tau fibrillization by blocking filament
extension: implications for the treatment of tauopathic
neurodegenerative diseases. Biochemistry 44,
10227–10237 (2005).
67.
Pickhardt, M. et al. Screening for inhibitors of tau
polymerization. Curr. Alzheimer Res. 2, 219–226
(2005).
68.
Pickhardt, M. et al. Anthraquinones inhibit tau
aggregation and dissolve Alzheimer’s paired helical
filaments in vitro and in cells. J. Biol. Chem. 280,
3628–3635 (2005).
69.
Taniguchi, S. et al. Inhibition of heparin-induced tau
filament formation by phenothiazines, polyphenols,
and porphyrins. J. Biol. Chem. 280, 7614–7623
(2005).
70.
Frid, P., Anisimov, S. V. & Popovic, N. Congo red and
protein aggregation in neurodegenerative diseases.
Brain Res. Rev. 53, 135–160 (2007).
71.
Chirita, C., Necula, M. & Kuret, J. Ligand-dependent
inhibition and reversal of tau filament formation.
Biochemistry 43, 2879–2887 (2004).
72.
Liu, M., Ni, J., Kosik, K. S. & Yeh, L. A. Development of
a fluorescent high throughput assay for tau
aggregation. Assay Drug Dev. Technol. 2, 609–619
(2004).
73.
Wischik, C. M., Edwards, P. C., Lai, R. Y. K., Roth, M. &
Harrington, C. R. Selective inhibition of Alzheimer
disease-like tau aggregation by phenothiazines.
Proc. Natl Acad. Sci. USA 93, 11213–11218
(1996).
74.
Ignatova, Z. & Gierasch, L. M. Inhibition of protein
aggregation in vitro and in vivo by a natural
osmoprotectant. Proc. Natl Acad. Sci. USA 103,
13357–13361 (2006).
75.
Gotz, J. et al. A decade of tau transgenic animal
models and beyond. Brain Pathol. 17, 91–103
(2007).
An up‑to‑date account of tau transgenic animal
models.
76.
McGowan, E., Eriksen, J. & Hutton, M. A decade of
modeling Alzheimer’s disease in transgenic mice.
Trends Genet. 22, 281–289 (2006).
77.
Van Dam, D. & De Deyn, P. P. Drug discovery in
dementia: the role of rodent models. Nature Rev. Drug
Discov. 5, 956–970 (2006).
78.
Schindowski, K. et al. Alzheimer’s disease-like tau
neuropathology leads to memory deficits and loss of
functional synapses in a novel mutated tau transgenic
mouse without any motor deficits. Am. J. Pathol. 169,
599–616 (2006).
79.
Melnikova, I. Therapies for Alzheimer’s disease.
Nature Rev. Drug Discov. 6, 341–342 (2007).
80.
Okamura, N. et al. Quinoline and benzimidazole
derivatives: candidate probes for in vivo imaging of
tau pathology in Alzheimer’s disease. J. Neurosci. 25,
10857–10862 (2005).
81.
Pardridge, W. M. The blood–brain barrier:
bottleneck in brain drug development. NeuroRx 2,
3–14 (2005).
82.
Gotz, J. et al. Somatodendritic localization and
hyperphosphorylation of tau protein in transgenic
mice expressing the longest human brain tau isoform.
EMBO J. 14, 1304–1313 (1995).
83.
Spittaels, K. et al. Prominent axonopathy in the brain
and spinal cord of transgenic mice overexpressing
four-repeat human tau protein. Am. J. Pathol. 155,
2153–2165 (1999).
84.
Probst, A. et al. Axonopathy and amyotrophy in mice
transgenic for human four-repeat tau protein. Acta
Neuropathol. 99, 469–481 (2000).
85.
Ishihara, T. et al. Age-dependent induction of
congophilic neurofibrillary tau inclusions in tau
transgenic mice. Am. J. Pathol. 158, 555–562
(2001).
86.
Lewis, J. et al. Neurofibrillary tangles, amyotrophy
and progressive motor disturbance in mice expressing
mutant (P301L) tau protein. Nature Genet. 25,
402–405 (2000).
87.
Gotz, J., Chen, F., Barmettler, R. & Nitsch, R. M.
Tau filament formation in transgenic mice expressing
P301L tau. J. Biol. Chem. 276, 529–534 (2001).
88.
Duff, K. et al. Characterization of pathology in
transgenic mice over-expressing human genomic and
cDNA tau transgenes. Neurobiol. Dis. 7, 87–98
(2000).
89.
Gotz, J. et al. Oligodendroglial tau filament formation
in transgenic mice expressing G272V tau. Eur. J.
Neurosci. 13, 2131–2140 (2001).
90.
Allen, B. et al. Abundant tau filaments and
nonapoptotic neurodegeneration in transgenic mice
expressing human P301S tau protein. J. Neurosci. 22,
9340–9351 (2002).
91.
Lim, F. et al. FTDP-17 mutations in tau transgenic
mice provoke lysosomal abnormalities and tau
filaments in forebrain. Mol. Cell Neurosci. 18,
702–714 (2001).
92.
Tanemura, K. et al. Neurodegeneration with tau
accumulation in a transgenic mouse expressing
V337M human tau. J. Neurosci. 22, 133–141
(2002).
93.
Tatebayashi, Y. et al. Tau filament formation and
associative memory deficit in aged mice expressing
mutant (R406W) human tau. Proc. Natl Acad. Sci.
USA 99, 13896–13901 (2002).
94.
Tesseur, I. et al. Prominent axonopathy and
disruption of axonal transport in transgenic mice
expressing human apolipoprotein E4 in neurons of
brain and spinal cord. Am. J. Pathol. 157,
1495–1510 (2000).
95.
Ahlijanian, M. K. et al. Hyperphosphorylated tau
and neurofilament and cytoskeletal disruptions in
mice overexpressing human p25, an activator of
cdk5. Proc. Natl Acad. Sci. USA 97, 2910–2915
(2000).
REVIEWS
NATURE REVIEWS
|
NEUROSCIENCE VOLUME 8
|
SEPTEMBER 2007
|
671
© 2007 Nature Publishing Group
96. Capsoni, S. et al. Alzheimer-like neurodegeneration in
aged antinerve growth factor transgenic mice. Proc.
Natl Acad. Sci. USA 97, 6826–6831 (2000).
97.
Oddo, S., Caccamo, A., Kitazawa, M., Tseng, B. P. &
LaFerla, F. M. Amyloid deposition precedes tangle
formation in a triple transgenic model of Alzheimer’s
disease. Neurobiol. Aging 24, 1063–1070 (2003).
98.
Oddo, S. et al. Triple-transgenic model of Alzheimer’s
disease with plaques and tangles: intracellular Aβ and
synaptic dysfunction. Neuron 39, 409–421 (2003).
99.
Crowe, A., Ballatore, C., Hyde, E., Trojanowski, J. Q. &
Lee, V. M.-Y. High throughput screening for small
molecule inhibitors of heparin-induced tau fibril
formation. Biochem. Biophys. Res. Commun. 358,
1–6 (2007).
100.
Dickey, C. A. et al. The high-affinity HSP90-CHIP
complex recognizes and selectively degrades
phosphorylated tau client proteins. J. Clin. Invest.
117, 648–658 (2007).
101.
Goryunov, D. & Liem, R. K. H. CHIP-ping away at tau.
J. Clin. Invest. 117, 590–592 (2007).
102.
Matsuoka, Y. et al. Intranasal NAP administration
reduces accumulation of amyloid peptide and tau
hyperphosphorylation in a transgenic mouse
model of Alzheimer’s disease at early pathological
stage. J. Mol. Neurosci. 31, 165–170 (2007).
103.
Pasinetti, G. M. From epidemiology to therapeutic
trials with anti-inflammatory drugs in Alzheimer’s
disease: the role of NSAIDs and cyclooxygenase in β-
amyloidosis and clinical dementia. J. Alzheimers Dis.
4, 435–445 (2002).
104.
Klegeris, A. & McGeer, P. L. Non-steroidal anti-
inflammatory drugs (NSAIDs) and other anti-inflammatory
agents in the treatment of neurodegenerative disease.
Curr. Alzheimer Res. 2, 355–365 (2005).
105.
Townsend, K. P. & Pratico, D. Novel therapeutic
opportunities for Alzheimer’s disease: focus on
nonsteroidal anti-inflammatory drugs. FASEB J. 19,
1592–1601 (2005).
Acknowledgements
We thank our colleagues for their contributions to the work
summarized here, which has been supported by grants from
the US National Institutes of Health (P01 AG09215, P30
AG10124, P01 AG11542, P01 AG14382, P01 AG14449,
P01 AG17586, PO1 AG19724, P01 NS-044233, UO1
AG24904), and the Marian S. Ware Alzheimer Program.
Finally, we are indebted to our patients and their families,
whose commitment to research has made our work possible.
Competing interests statement
The authors declare no competing financial interests.
DATABASES
Entrez Gene: http://www.ncbi.nlm.nih.gov/entrez/query.
fcgi?db=gene
APP | CDK5 | FYN | GSK3α | GSK3β | MAPT | MARK | presenilin 1
OMIM: http://www.ncbi.nlm.nih.gov/entrez/query.
fcgi?db=OMIM
Alzheimer’s disease
FURTHER INFORMATION
John Q. Trojanowski’s homepages:
http://www.med.upenn.edu/aging
http://www.uphs.upenn.edu/ADC
http://www.uphs.upenn.edu/cndr/
ALL LINKS ARE ACTIVE IN THE ONLINE PDF.
REVIEWS
672
|
SEPTEMBER 2007
|
VOLUME 8 www.nature.com/reviews/neuro
... Upon TBI, the high strain placed on axons may inhibit the mobility of adjacent microtubules to slide past each other in response to axonal stretching, resulting in long microtubule breakage and short microtubule detachment from microtubule bundles [76]. These mechanical forces at the site of injury promote soluble tau detachment from the microtubule [77,78], facilitating post-translational modifications at the disease-associated residues and increasing the propensity of tau to aggregate into oligomers and NFTs [79,80]. While some studies have shown that repeated TBI (such as in CTE) increases tau hyperphosphorylation [81], fairly little is known about changes in the abundance of other forms of pathological tau species upon a single mild injury. ...
Article
Full-text available
Background Tau is aberrantly acetylated in various neurodegenerative conditions, including Alzheimer’s disease, frontotemporal lobar degeneration (FTLD), and traumatic brain injury (TBI). Previously, we reported that reducing acetylated tau by pharmacologically inhibiting p300-mediated tau acetylation at lysine 174 reduces tau pathology and improves cognitive function in animal models. Methods We investigated the therapeutic efficacy of two different antibodies that specifically target acetylated lysine 174 on tau (ac-tauK174). We treated PS19 mice, which harbor the P301S tauopathy mutation that causes FTLD, with anti-ac-tauK174 and measured effects on tau pathology, neurodegeneration, and neurobehavioral outcomes. Furthermore, PS19 mice received treatment post-TBI to evaluate the ability of the immunotherapy to prevent TBI-induced exacerbation of tauopathy phenotypes. Ac-tauK174 measurements in human plasma following TBI were also collected to establish a link between trauma and acetylated tau levels, and single nuclei RNA-sequencing of post-TBI brain tissues from treated mice provided insights into the molecular mechanisms underlying the observed treatment effects. Results Anti-ac-tauK174 treatment mitigates neurobehavioral impairment and reduces tau pathology in PS19 mice. Ac-tauK174 increases significantly in human plasma 24 h after TBI, and anti-ac-tauK174 treatment of PS19 mice blocked TBI-induced neurodegeneration and preserved memory functions. Anti-ac-tauK174 treatment rescues alterations of microglial and oligodendrocyte transcriptomic states following TBI in PS19 mice. Conclusions The ability of anti-ac-tauK174 treatment to rescue neurobehavioral impairment, reduce tau pathology, and rescue glial responses demonstrates that targeting tau acetylation at K174 is a promising neuroprotective therapeutic approach to human tauopathies resulting from TBI or genetic disease.
... Neurodegenerative tauopathies, which include Alzheimer's disease (AD), are defined by the presence of intracellular inclusions composed of tau protein within brain neurons. 1,2 Tau inclusions are the primary neuropathological feature in non-AD tauopathies, which include frontotemporal lobar degeneration (FTLD), progressive supranuclear palsy (PSP), corticobasal degeneration (CBD), Pick's disease, chronic traumatic encephalopathy, and primary age-related tauopathy. [2][3][4][5] In contrast, those with AD also have senile plaques composed of amyloid beta (Aβ) peptides. ...
Article
Full-text available
INTRODUCTION Intraneuronal inclusions composed of tau protein are found in Alzheimer's disease (AD) and other tauopathies. Tau normally binds microtubules (MTs), and its disengagement from MTs and misfolding in AD is thought to result in MT abnormalities. We previously identified triazolopyrimidine‐containing MT‐stabilizing compounds that provided benefit in AD mouse models and herein describe the characterization and efficacy testing of an optimized candidate, CNDR‐51997. METHODS CNDR‐51997 underwent pharmacokinetic, pharmacodynamic, safety pharmacology, and mouse tolerability testing. In addition, the compound was examined for efficacy in 5XFAD amyloid beta (Aβ) plaque mice and PS19 tauopathy mice. RESULTS CNDR‐51997 significantly reduced Aβ plaques in 5XFAD mice and tau pathology in PS19 mice, with the latter also showing attenuated axonal dystrophy and gliosis. CNDR‐51997 was well tolerated at doses that exceeded efficacy doses, with a good safety pharmacology profile. DISCUSSION CNDR‐51997 may be a candidate for advancement as a potential therapeutic agent for AD and/or other tauopathies. Highlights There is evidence of microtubule alterations (MT) in Alzheimer's disease (AD) brain and in mouse models of AD pathology. Intermittent dosing with an optimized, brain‐penetrant MT‐stabilizing small‐molecule, CNDR‐51997, reduced both Aβ plaque and tau inclusion pathology in established mouse models of AD. CNDR‐51997 attenuated axonal dystrophy and gliosis in a tauopathy mouse model, with a strong trend toward reduced hippocampal neuron loss. CNDR‐51997 is well tolerated in mice at doses that are meaningfully greater than required for efficacy in AD mouse models, and the compound has a good safety pharmacology profile.
... Since the proposal of AD by Alois Alzheimer in 1906, through the pathological anatomy of a woman, the exact pathogenesis of AD had not yet been definitively identified. However, various hypotheses have been put forward, most notably the amyloid However, various hypotheses have been put forward, most notably the amyloid cascade hypothesis, Tau neurofibrillary tangles hypothesis, and neuroinflammation hypothesis [12][13][14][15][16]. With progressive research, it has also become evident that there is a certain level of interaction between these hypotheses [17,18]. ...
Article
Full-text available
As the population ages worldwide, Alzheimer’s disease (AD), the most prevalent kind of neurodegenerative disorder among older people, has become a significant factor affecting quality of life, public health, and economies. However, the exact pathogenesis of Alzheimer’s remains elusive, and existing highly recognized pathogenesis includes the amyloid cascade hypothesis, Tau neurofibrillary tangles hypothesis, and neuroinflammation hypothesis. The major diagnoses of Alzheimer’s disease include neuroimaging positron emission computed tomography, magnetic resonance imaging, and cerebrospinal fluid molecular diagnosis. The therapy of Alzheimer’s disease primarily relies on drugs, and the approved drugs on the market include acetylcholinesterase drugs, glutamate receptor antagonists, and amyloid-β monoclonal antibodies. Still, the existing drugs can only alleviate the symptoms of the disease and cannot completely reverse it. This review aims to summarize existing research results on Alzheimer’s disease pathogenesis, diagnosis, and drug therapy, with the objective of facilitating future research in this area.
... Tauopathies represent a diverse group of over twenty neurodegenerative diseases characterized by hyperphosphorylated and aggregated Tau forms in neuronal and glial cells (Table 1) [15,16]. These diseases exhibit varying clinical phenotypes and pathophysiological characteristics [17]. ...
Article
Full-text available
Neurological damage is the pathological substrate of permanent disability in various neurodegenerative disorders. Early detection of this damage, including its identification and quantification, is critical to preventing the disease’s progression in the brain. Tau, glial fibrillary acidic protein (GFAP), and neurofilament light chain (NfL), as brain protein biomarkers, have the potential to improve diagnostic accuracy, disease monitoring, prognostic assessment, and treatment efficacy. These biomarkers are released into the cerebrospinal fluid (CSF) and blood proportionally to the degree of neuron and astrocyte damage in different neurological disorders, including stroke, traumatic brain injury, multiple sclerosis, neurodegenerative dementia, and Parkinson’s disease. Here, we review how Tau, GFAP, and NfL biomarkers are detected in CSF and blood as crucial diagnostic tools, as well as the levels of these biomarkers used for differentiating a range of neurological diseases and monitoring disease progression. We also discuss a biosensor approach that allows for the real-time detection of multiple biomarkers in various neurodegenerative diseases. This combined detection system of brain protein biomarkers holds significant promise for developing more specific and accurate clinical tools that can identify the type and stage of human neurological diseases with greater precision.
... Two phosphothreonine-containing peptides derived from Tau (ref. 33) and a phosphoserine-containing peptide derived from TAK1 (ref. 34) were further tested as substrates for LnaB in vitro (Extended Data Figs. ...
Article
Full-text available
AMPylation is a post-translational modification in which AMP is added to the amino acid side chains of proteins1,2. Here we show that, with ATP as the ligand and actin as the host activator, the effector protein LnaB of Legionella pneumophila exhibits AMPylase activity towards the phosphoryl group of phosphoribose on PRR42-Ub that is generated by the SidE family of effectors, and deubiquitinases DupA and DupB in an E1- and E2-independent ubiquitination process3–7. The product of LnaB is further hydrolysed by an ADP-ribosylhydrolase, MavL, to Ub, thereby preventing the accumulation of PRR42-Ub and ADPRR42-Ub and protecting canonical ubiquitination in host cells. LnaB represents a large family of AMPylases that adopt a common structural fold, distinct from those of the previously known AMPylases, and LnaB homologues are found in more than 20 species of bacterial pathogens. Moreover, LnaB also exhibits robust phosphoryl AMPylase activity towards phosphorylated residues and produces unique ADPylation modifications in proteins. During infection, LnaB AMPylates the conserved phosphorylated tyrosine residues in the activation loop of the Src family of kinases8,9, which dampens downstream phosphorylation signalling in the host. Structural studies reveal the actin-dependent activation and catalytic mechanisms of the LnaB family of AMPylases. This study identifies, to our knowledge, an unprecedented molecular regulation mechanism in bacterial pathogenesis and protein phosphorylation.
Article
Introduction There are limited data regarding cerebrospinal fluid (CSF) and plasma biomarkers among patients with Cerebral Amyloid Angiopathy (CAA). We sought to investigate the levels of four biomarkers [β-amyloids (Aβ42 and Aβ40), total tau (tau) and phosphorylated tau (p-tau)] in CAA patients compared to healthy controls (HC) and patients with Alzheimer Disease (AD). Patients and methods A systematic review and meta-analysis of published studies, including also a 5 year single-center cohort study, with available data on CSF and plasma biomarkers in symptomatic sporadic CAA versus HC and AD was conducted. Biomarkers’ comparisons were investigated using random-effects models based on the ratio of mean (RoM) biomarker concentrations. RoM < 1 and RoM > 1 indicate lower and higher biomarker concentration in CAA compared to another population, respectively. Results We identified nine cohorts, comprising 327 CAA patients (mean age: 71 ± 5 years; women: 45%) versus 336 HC (mean age: 65 ± 5 years; women: 45%) and 384 AD patients (mean age: 68 ± 3 years; women: 53%) with available data on CSF biomarkers. CSF Aβ42 levels [RoM: 0.47; 95% CI: 0.36–0.62; p < 0.0001], Aβ40 levels [RoM: 0.70; 95% CI: 0.63–0.79; p < 0.0001] and the ratio Aβ42/Aβ40 [RoM: 0.62; 95% CI: 0.39–0.98; p = 0.0438] differentiated CAA from HC. CSF Aβ40 levels [RoM: 0.73; 95% CI: 0.64–0.83; p = 0.0003] differentiated CAA from AD. CSF tau and p-tau levels differentiated CAA from HC [RoM: 1.71; 95% CI: 1.41–2.09; p = 0.0002 and RoM: 1.44; 95% CI: 1.20–1.73; p = 0.0014, respectively] and from AD [RoM: 0.65; 95% CI: 0.58–0.72; p < 0.0001 and RoM: 0.64; 95% CI: 0.57–0.71; p < 0.0001, respectively]. Plasma Aβ42 [RoM: 1.14; 95% CI: 0.89–1.45; p = 0.2079] and Aβ40 [RoM: 1.07; 95% CI: 0.91–1.25; p = 0.3306] levels were comparable between CAA and HC. Conclusions CAA is characterized by a distinct CSF biomarker pattern compared to HC and AD. CSF Aβ40 levels are lower in CAA compared to HC and AD, while tau and p-tau levels are higher in CAA compared to HC, but lower in comparison to AD patients.
Preprint
Full-text available
Proteomic profiling of Alzheimer's disease (AD) brains has identified numerous understudied proteins, including midkine (MDK), that are highly upregulated and correlated with Aβ since the early disease stage, but their roles in disease progression are not fully understood. Here we present that MDK attenuates Aβ assembly and influences amyloid formation in the 5xFAD amyloidosis mouse model. MDK protein mitigates fibril formation of both Aβ40 and Aβ42 peptides in Thioflavin T fluorescence assay, circular dichroism, negative stain electron microscopy, and NMR analysis. Knockout of Mdk gene in 5xFAD increases amyloid formation and microglial activation. Further comprehensive mass spectrometry-based profiling of whole proteome and aggregated proteome in these mouse models indicates significant accumulation of Aβ and Aβ-correlated proteins, along with microglial components. Thus, our structural and mouse model studies reveal a protective role of MDK in counteracting amyloid pathology in Alzheimer's disease.
Article
Full-text available
GSK-3β, IKK-β, and ROCK-1 kinases are implicated in the pathomechanism of Alzheimer’s disease due to their involvement in the misfolding and accumulation of amyloid β (Aβ) and tau proteins, as well as inflammatory processes. Among these kinases, GSK-3β plays the most crucial role. In this study, we present compound 62, a novel, remarkably potent, competitive GSK-3β inhibitor (IC50 = 8 nM, Ki = 2 nM) that also exhibits additional ROCK-1 inhibitory activity (IC50 = 2.3 µM) and demonstrates anti-inflammatory and neuroprotective properties. Compound 62 effectively suppresses the production of nitric oxide (NO) and pro-inflammatory cytokines in the lipopolysaccharide-induced model of inflammation in the microglial BV-2 cell line. Furthermore, it shows neuroprotective effects in an okadaic-acid-induced tau hyperphosphorylation cell model of neurodegeneration. The compound also demonstrates the potential for further development, characterized by its chemical and metabolic stability in mouse microsomes and fair solubility.
Article
Full-text available
The neuropathological correlates of Alzheimer's disease (AD) include amyloid-β (Aβ) plaques and neurofibrillary tangles. To study the interaction between Aβ and tau and their effect on synaptic function, we derived a triple-transgenic model (3×Tg-AD) harboring PS1M146V, APPSwe, and tauP301L transgenes. Rather than crossing independent lines, we microinjected two transgenes into single-cell embryos from homozygous PS1M146V knockin mice, generating mice with the same genetic background. 3×Tg-AD mice progressively develop plaques and tangles. Synaptic dysfunction, including LTP deficits, manifests in an age-related manner, but before plaque and tangle pathology. Deficits in long-term synaptic plasticity correlate with the accumulation of intraneuronal Aβ. These studies suggest a novel pathogenic role for intraneuronal Aβ with regards to synaptic plasticity. The recapitulation of salient features of AD in these mice clarifies the relationships between Aβ, synaptic dysfunction, and tangles and provides a valuable model for evaluating potential AD therapeutics as the impact on both lesions can be assessed.
Article
Full-text available
Families bearing mutations in the presenilin 1 (PS1) gene develop Alzheimer’s disease. Previous studies have shown that the Alzheimer-associated mutations in PS1 increase production of amyloid β protein (Aβ1–42). We now show that PS1 also regulates phosphorylation of the microtubule-associated protein tau. PS1 directly binds tau and a tau kinase, glycogen synthase kinase 3β (GSK-3β). Deletion studies show that both tau and GSK-3β bind to the same region of PS1, residues 250–298, whereas the binding domain on tau is the microtubule-binding repeat region. The ability of PS1 to bring tau and GSK-3β into close proximity suggests that PS1 may regulate the interaction of tau with GSK-3β. Mutations in PS1 that cause Alzheimer’s disease increase the ability of PS1 to bind GSK-3β and, correspondingly, increase its tau-directed kinase activity. We propose that the increased association of GSK-3β with mutant PS1 leads to increased phosphorylation of tau.
Article
Full-text available
Alzheimer's disease is associated with increased production and aggregation of amyloid-β (Aβ) peptides. Aβ peptides are derived from the amyloid precursor protein (APP) by sequential proteolysis, catalysed by the aspartyl protease BACE, followed by presenilin-dependent γ-secretase cleavage. Presenilin interacts with nicastrin, APH-1 and PEN-2 (ref. 6), all of which are required for γ-secretase function. Presenilins also interact with α-catenin, β-catenin and glycogen synthase kinase-3β (GSK-3β), but a functional role for these proteins in γ-secretase activity has not been established. Here we show that therapeutic concentrations of lithium, a GSK-3 inhibitor, block the production of Aβ peptides by interfering with APP cleavage at the γ-secretase step, but do not inhibit Notch processing. Importantly, lithium also blocks the accumulation of Aβ peptides in the brains of mice that overproduce APP. The target of lithium in this setting is GSK-3α, which is required for maximal processing of APP. Since GSK-3 also phosphorylates tau protein, the principal component of neurofibrillary tangles, inhibition of GSK-3α offers a new approach to reduce the formation of both amyloid plaques and neurofibrillary tangles, two pathological hallmarks of Alzheimer's disease.
Article
We have determined the biochemical and immunocytochemical localization of the heterogeneous microtubule-associated protein tau using a monoclonal antibody that binds to all of the tau polypeptides in both bovine and rat brain. Using immunoblot assays and competitive enzyme-linked immunosorbent assays, we have shown tau to be more abundant in bovine white matter extracts and microtubules than in extracts and microtubules from an enriched gray matter region of the brain. On a per mole basis, twice-cycled microtubules from white matter contained three times more tau than did twice-cycled microtubules from gray matter. Immunohistochemical studies that compared the localization of tau with that of MAP2 and tubulin demonstrated that tau was restricted to axons, extending the results of the biochemical studies. Tau localization was not observed in glia, which indicated that, at least in brain, tau is neuron specific. These observations indicate that tau may help define a subpopulation of microtubules that is restricted to axons. Furthermore, the monoclonal antibody described in this report should prove very useful to investigators studying axonal sprouting and growth because it is an exclusive axonal marker.
Article
Amyloid-β (Aβ) containing plaques and tau-laden neurofibrillary tangles are the defining neuropathological features of Alzheimer’s disease (AD). To better mimic this neuropathology, we generated a novel triple transgenic model of AD (3xTg-AD) harboring three mutant genes: β-amyloid precursor protein (βAPPSwe), presenilin-1 (PS1M146V), and tauP301L. The 3xTg-AD mice progressively develop Aβ and tau pathology, with a temporal- and regional-specific profile that closely mimics their development in the human AD brain. We find that Aβ deposits initiate in the cortex and progress to the hippocampus with aging, whereas tau pathology is first apparent in the hippocampus and then progresses to the cortex. Despite equivalent overexpression of the human βAPP and human tau transgenes, Aβ deposition develops prior to the tangle pathology, consistent with the amyloid cascade hypothesis. As these 3xTg-AD mice phenocopy critical aspects of AD neuropathology, this model will be useful in pre-clinical intervention trials, particularly because the efficacy of anti-AD compounds in mitigating the neurodegenerative effects mediated by both signature lesions can be evaluated.
Article
Since the initial description one hundred years ago by Dr. Alois Alzheimer, the disorder that bears his name has been characterized by the occurrence of two brain lesions: amyloid plaques and neurofibrillary tangles (NFTs). Yet the precise relationship between beta-amyloid (Aβ) and tau, the two proteins that accumulate within these lesions, has proven elusive. Today, a growing body of work supports the notion that Aβ may directly or indirectly interact with tau to accelerate NFT formation. Here we review recent evidence that Aβ can adversely affect distinct molecular and cellular pathways, thereby facilitating tau phosphorylation, aggregation, mis-localization, and accumulation. Studies are presented that support four putative mechanisms by which Aβ may facilitate the development of tau pathology. A great deal of work suggests that Aβ may drive tau pathology by activating specific kinases, providing a straightforward mechanism by which Aβ may enhance tau hyperphosphorylation and NFT formation. In the AD brain, Aβ also triggers a massive inflammatory response and pro-inflammatory cytokines can in turn indirectly modulate tau phosphorylation. Mounting evidence also suggests that Aβ may inhibit tau degradation via the proteasome. Lastly, Aβ and tau may indirectly interact at the level of axonal transport and evidence is presented for two possible scenarios by which axonal transport deficits may play a role. We propose that the four putative mechanisms described in this review likely mediate the interactions between Aβ and tau, thereby leading to the development of AD neurodegeneration.
Article
For an increasing number of neurodegenerative diseases, like Alzheimer’s disease, Parkinson’s disease and polyglutamine disease, e.g. Huntington’s disease, the aggregation of misfolded protein is thought to play a crucial role in the neuropathy. Recently, a new gene has been identified in a Caenorhabditis elegans (C.elegans) model for polyglutamine aggregation, called modifier of aggregation 4 (moag-4). Inactivation of moag-4 suppressed aggregation by 75% and reduced polyglutamine toxicity. These effects have been replicated in C.elegans models for aggregation in Alzheimer’s disease and Parkinson’s disease. Moreover, moag-4 has two human orthologues, Serf1a and Serf2, which in cell culture show a similar effect on aggregation and toxicity of mutant huntingtin. It is unknown how moag-4 influences aggregation. We hypothesized that moag-4 acts on aggregation through downstream targets. In order to investigate this, we performed a RNA interference screen, in which we knocked-down selected genes that have been previously associated with protein aggregation and determined whether they could reverse the effect of moag-4 inactivation, what would indicate that moag-4 acts through these genes. We included genes coding for transcription factors of the insulin/IGF-1 signalling pathway, heat shock proteins, components of CCT chaperonin, predicted chaperones, the ubiquitin-proteasome system and C.elegans orthologues of genes associated to aggregation in other species. Knock-down of none of the included genes could rescue the suppression of aggregation by loss-of-function of moag-4. These findings strongly suggest that moag-4 acts independently of genes known to influence protein aggregation in C.elegans and therefore might represent a novel pathway in protein folding and neurodegenerative diseases.
Article
The epsilon 4 allele of the human apolipoprotein E gene (ApoE4) constitutes an important genetic risk factor for Alzheimer’s disease. Recent experimental evidence suggests that human ApoE is expressed in neurons, in addition to being synthesized in glial cells. Moreover, brain regions in which neurons express ApoE seem to be most vulnerable to neurofibrillary pathology. The hypothesis that the expression pattern of human ApoE might be important for the pathogenesis of Alzheimer’s disease was tested by generating transgenic mice that express human ApoE4 in neurons or in astrocytes of the central nervous system. Transgenic mice expressing human ApoE4 in neurons developed axonal degeneration and gliosis in brain and in spinal cord, resulting in reduced sensorimotor capacities. In these mice, axonal dilatations with accumulation of synaptophysin, neurofilaments, mitochondria, and vesicles were documented, suggesting impairment of axonal transport. In contrast, transgenic mice expressing human ApoE4 in astrocytes remained normal throughout life. These results suggest that expression of human ApoE in neurons of the central nervous system could contribute to impaired axonal transport and axonal degeneration. The possible contribution of hyperphosphorylation of protein Tau to the resulting phenotype is discussed.