ArticlePDF AvailableLiterature Review

Abstract and Figures

Major discoveries have been made in the recent past in the genetics, biochemistry and neuropathology of frontotemporal lobar degeneration (FTLD). TAR DNA-binding protein 43 (TDP-43), encoded by the TARDBP gene, has been identified as the major pathological protein of FTLD with ubiquitin-immunoreactive (ub-ir) inclusions (FTLD-U) with or without amyotrophic lateral sclerosis (ALS) and sporadic ALS. Recently, mutations in the TARDBP gene in familial and sporadic ALS have been reported which demonstrate that abnormal TDP-43 alone is sufficient to cause neurodegeneration. Several familial cases of FTLD-U, however, are now known to have mutations in the progranulin (GRN) gene, but granulin is not a component of the TDP-43- and ub-ir inclusions. Further, TDP-43 is found to be a component of the inclusions of an increasing number of neurodegenerative diseases. Other FTLD-U entities with TDP-43 proteinopathy include: FTLD-U with valosin-containing protein (VCP) gene mutation and FTLD with ALS linked to chromosome 9p. In contrast, chromosome 3-linked dementia, FTLD-U with chromatin modifying protein 2B (CHMP2B) mutation, has ub-ir, TDP-43-negative inclusions. In summary, recent discoveries have generated new insights into the pathogenesis of a spectrum of disorders called TDP-43 proteinopathies including: FTLD-U, FTLD-U with ALS, ALS, and a broadening spectrum of other disorders. It is anticipated that these discoveries and a revised nosology of FTLD will contribute toward an accurate diagnosis, and facilitate the development of new diagnostic tests and therapeutics.
Content may be subject to copyright.
ALS and FTLD: two faces of TDP-43 proteinopathy
R. M. Liscica,b, L. T. Grinbergb,c,d,e, J. Zidarf, M. A. Gitchob,g, and N. J. Cairnsb,e,g
a Institute for Medical Research and Occupational Health, Zagreb, Croatia
b Alzheimer’s Disease Research Center, Washington University School of Medicine, St Louis,
MO, USA
c Department of Pathology, University of São Paulo Medical School, São Paulo, Brazil
d Instituto Israelita de Ensino e Pesquisa Albert Einstein, São Paulo, Brazil
e Pathology & Immunology, Washington University School of Medicine, St Louis, MO, USA
f Institute of Clinical Neurophysiology, University Medical Centre, Ljubljana, Slovenia
g Department of Neurology, Washington University School of Medicine, St Louis, MO, USA
Abstract
Major discoveries have been made in the recent past in the genetics, biochemistry and
neuropathology of frontotemporal lobar degeneration (FTLD). TAR DNA-binding protein 43
(TDP-43), encoded by the TARDBP gene, has been identified as the major pathological protein of
FTLD with ubiquitin-immunoreactive (ub-ir) inclusions (FTLD-U) with or without amyotrophic
lateral sclerosis (ALS) and sporadic ALS. Recently, mutations in the TARDBP gene in familial
and sporadic ALS have been reported which demonstrate that abnormal TDP-43 alone is sufficient
to cause neurodegeneration. Several familial cases of FTLD-U, however, are now known to have
mutations in the progranulin (GRN) gene, but granulin is not a component of the TDP-43- and ub-
ir inclusions. Further, TDP-43 is found to be a component of the inclusions of an increasing
number of neurodegenerative diseases. Other FTLD-U entities with TDP-43 proteinopathy
include: FTLD-U with valosin-containing protein (VCP) gene mutation and FTLD with ALS
linked to chromosome 9p. In contrast, chromosome 3-linked dementia, FTLD-U with chromatin
modifying protein 2B (CHMP2B) mutation, has ub-ir, TDP-43-negative inclusions. In summary,
recent discoveries have generated new insights into the pathogenesis of a spectrum of disorders
called TDP-43 proteinopathies including: FTLD-U, FTLD-U with ALS, ALS, and a broadening
spectrum of other disorders. It is anticipated that these discoveries and a revised nosology of
FTLD will contribute toward an accurate diagnosis, and facilitate the development of new
diagnostic tests and therapeutics.
Keywords
amyotrophic lateral sclerosis; frontotemporal dementia; frontotemporal lobar degeneration;
granulin; motor neuron disease; TARDBP; TDP-43; ubiquitin; valosin-containing protein
Correspondence: Nigel J. Cairns, PhD, FRCPath, Department of Pathology and Immunology, Washington University School of
Medicine, Campus Box 8118, 660 South Euclid Avenue, St Louis, MO 63110, USA (tel.: +1 314 362 7420; fax: +1 314 362 4096;
cairns@wustl.edu).
Disclosure
The authors report no conflicts of interest.
NIH Public Access
Author Manuscript
Eur J Neurol. Author manuscript; available in PMC 2010 January 4.
Published in final edited form as:
Eur J Neurol
. 2008 August ; 15(8): 772–780. doi:10.1111/j.1468-1331.2008.02195.x.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
Introduction
Frontotemporal lobar degeneration (FTLD) is used here as an umbrella term to include both
a clinical syndrome and one of the neuropathological entities [1–3]. FTLD is a focal, non-
Alzheimer form of dementia, clinically characterized as either behavioral or aphasic variants
[1,2,4]. Most commonly, the behavioral or frontal variant is characterized by behavioral
dysfunction and change in personal and social conduct. The aphasic variant includes a non-
fluent form called progressive non-fluent aphasia, and a fluent form called semantic
dementia. Typically, the patient with FTLD does not have an amnestic syndrome, at least in
the early stage of the disease, which distinguishes FTLD clinically from Alzheimer’s disease
(AD) [5], but there are exceptions [6]. Focal dementias account for up to 20% of presenile
dementia cases [7], and FTLD is the second most frequent form of dementia in people under
the age of 65 years after AD [8]. FTLD may occur alone or in combination with
amyotrophic lateral sclerosis (ALS), parkinsonism, or corticobasal syndrome during the
course of the disease.
ALS is the most common adult-onset progressive and, ultimately fatal, motor neuron disease
(MND). Like FTLD, ALS encompasses a range of clinicopathological entities [9]. ALS is
used here to describe signs of upper and lower motor neuron degeneration with a
progressive spread of signs within a region or to other regions as defined by the El Escorial
World Federation of Neurology Criteria [10]. The overlap between dementia and ALS is
demonstrated by the presence of cognitive and behavioral dysfunction in up to 50% of ALS
patients [10–16], indicating a spectrum of clinical phenotypes that relate to common
neuropathological lesions [17–19]. Further evidence for a clinical overlap between FTLD
and ALS is the occurrence of progressive aphasia [20,21] and the presence of
frontotemporal atrophy [22] in patients with ALS. Incidence rates of FTLD with ALS vary
often as a consequence of referral bias and differing diagnostic criteria. Several studies have
shown that ALS patients with frontotemporal impairment have significantly shorter survival
compared with FTLD patients [23,24]. However, the clinical diagnosis of FTLD may only
be considered after other potential causes of dementia (e.g. small and/or large vessel
disease), systemic conditions (e.g. hypothyroidism, B-vitamin deficiency), tumors, and
substance abuse have been excluded. This review will highlight a number of important
advances in our understanding of the molecular genetics, biochemistry, and neuropathology
of FTLD that have occurred within the recent past.
Genetic studies
FTLD is a genetically complex disorder, with multiple genetic factors contributing to the
disease. A positive family history with an autosomal dominant pattern of inheritance and
high penetrance is usually found in one quarter to one half of patients [25–30]. Recently,
several genes and a locus on chromosome 9p have been linked to familial FTLD with
ubiquitin-immunoreactive, tau-negative inclusions (FTLD-U): genetic defects include
mutations in the chromatin modifying protein 2B (CHMP2B gene), the cause of
chromosome 3-linked FTLD [31], and mutations in the valosin-containing protein (VCP)
gene, a cause of chromosome 9-linked FTLD [32,33]. Locus heterogeneity for FTLD and
ALS is indicated by the presence of other genetic loci at 9p [34,35]. Recently, the major
genetic cause of familial FTLD-U linked to chromosome 17 was identified as mutations in
the progranulin (GRN) [36,37] gene. This discovery was soon replicated by the
identification of other GRN mutations in the HDDD2 [38], PPA1 and PPA3 [39], and
HDDD1 [40] families. Although null mutations, mainly nonsense and frameshift mutations
resulting in premature stop codons, in GRN were the first to be identified and the causal
mechanism underlying FTLD-U, some of the families described have missense mutations
which are predicted to alter trafficking, protein folding, or processing leading to a GRN
Liscic et al. Page 2
Eur J Neurol. Author manuscript; available in PMC 2010 January 4.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
protein haploinsufficiency [36,37]. Thus, a spectrum of mutations in GRN leads to loss of
functional protein, the primary etiology of FTLD with GRN mutation.
This genetic heterogeneity is also reflected in clinical variability. FTLD with GRN mutation
has been described in families with corticobasal syndrome [41–43] and in familial and
sporadic patients from large FTLD cohorts [44–48]. In 37 patients with a single gene defect,
Arg493X, in 30 families with FTLD, there was a variable age at onset (range 44–69 years)
and clinical diagnoses included: frontotemporal dementia, primary progressive aphasia,
corticobasal syndrome, and Alzheimer’s disease [47]. In contrast, there is phenotypical
homogeneity in the PPA1 and PPA3 families, at least in the initial stage of disease [39].
Also, FTLD with GRN A9D, a missense mutation located in the signal peptide, has been
described in a family with corticobasal syndrome [43] as well as in another family
(HDDD2) with prominent behavioral and language dysfunction [38]. Thus, FTLD with GRN
mutation is both clinically, neuropathologically, and genetically heterogeneous [36–50].
Most of the mutations reported to date, but not all, lead to premature termination of the
coding sequence and nonsense mediated decay (NMD) resulting in functional loss of one
allele. The overall effect of these changes is likely to be a significant reduction in the growth
modulator activity of GRN. The various GRN mutations predict at least two disease
mechanisms – the partial loss of functional GRN (haploinsufficiency) [36,37] or functional
loss caused by mis-trafficking of mutant protein [38,49]. Further studies are required to
expand the spectrum of FTLD phenotypes with the several GRN mutations that have now
been reported [36–50]. (For the current list of pathogenic mutations in FTLD refer to:
http://www.molgen.ua.ac.be/FTDMutations/).
FTLD-U accounts for 5–15% of all dementia disorders [51]. Mutations in the microtubule-
associated protein tau (MAPT) gene on chromosome 17q21 which, coincidentally, is in
close proximity to the GRN gene, are known to be responsible for 10–20% of familial FTLD
[52]. The progranulin gene (GRN) is mutated in 5–10% of patients with FTLD and in about
20% of patients with familial FTLD [53], similar to that of FTLD with MAPT mutation [52].
The recent discovery of pathogenic missense mutations in a highly conserved glycine-rich,
heterogeneous ribonucleoprotein interacting domain of the TARDBP gene (Fig. 1) in
autosomal dominant ALS families and sporadic cases confirms the importance of TDP-43 in
the pathogenesis of ALS and demonstrates that defects in TARDBP are sufficient to cause
TDP-43 proteinopathy [54–57]. In none of these families, or in any of the sporadic cases,
was there evidence of FTLD. As additional families are identified, more heterogeneous
clinical phenotypes may emerge. As the glycine-rich domain (Fig. 1) is a ‘hot spot’ for the
mutations reported so far, dysregulation of mRNA splicing may be the functional
consequence of these gene defects. A great deal of work has yet to be done to elucidate the
normal function of TDP-43 and the cellular pathways disrupted by abnormal protein caused
by mutations in TARDBP. As some familial FTLD-U cases do not have GRN, VCP,
CHMP2B, or TARDBP mutations, it is likely that additional causal genes and genetic risk
factors for FTLD-U exist.
Neuropathology
FTLD comprises a neuropathologically heterogeneous group of neurodegenerative diseases,
which share the common feature of preferential degeneration of the frontal and temporal
lobes [1,4]. FTLD pathology can be broadly divided into two main classes, based on
abnormal accumulation of hyperphosphorylated tau protein: those with tau-immunoreactive
(tau-ir) neuronal and/or glial inclusions called tauopathies and those with ubiquitin-
immunoreactive (ub-ir), tau-negative inclusions, called FTLD-U. FTLD-U is the most
common entity within this group [58]. ALS is also accompanied by a wide range of
Liscic et al. Page 3
Eur J Neurol. Author manuscript; available in PMC 2010 January 4.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
neuropathological features in which both cortical (upper motor neuron), and either brainstem
motor neurons or anterior horn cells (lower motor neuron) are involved with a signature
lesion: abnormal accumulation of insoluble proteins, which are ubiquitinated, in the
cytoplasm of degenerating motor neurons [59]. Ub-ir inclusions are observed more
commonly in sporadic ALS (SALS), but they are also seen in familial ALS (FALS) with Cu/
Zn superoxide dismutase (SOD1) gene mutations [60,61]. Until recently, ubiquitin
immunohistochemistry (IHC) was the only method to detect abnormal protein aggregates in
FTLD-U, FTLD-U with ALS, and SALS. In this spectrum of diseases, the ubiquitinated
inclusions contain neither tau, nor α-synuclein, nor neuronal intermediate filament protein
epitopes.
Recently, TAR DNA-binding protein 43 (TDP-43) [18,19], was identified as the major
component of inclusions of sporadic and familial FTLD-U, with and without ALS, and
SALS (Fig. 2) [60,62]. However, the absence of pathological TDP-43 in familial cases
harboring SOD1 gene mutations implies that motor neuron degeneration may result from a
different mechanism in those cases [60]. It remains to be seen if TDP-43 may be useful in
differentiating SOD1-related ALS from SALS [60].
Neuropathologically, the TDP-43 proteinopathies are characterized by ubiquitin- and
TDP-43-ir neuronal cytoplasmic inclusions (NCIs), neuronal intranuclear inclusions (NIIs),
dystrophic neurites (DNs) and, in cases of MND, glial cytoplasmic inclusions (GCIs)
[18,19,63–67]. The pathological inclusions of these disorders may be distinguished from
other protein folding diseases which are characterized by abnormal aggregates of tau, α-
synuclein, β-amyloid, neuronal intermediate filament proteins, or expanded polyglutamine
repeats. TDP-43 proteinopathy is characterized biochemically by the presence of relatively
insoluble ubiquitinated TDP-43-containing inclusions of which TDP-43 is abnormally
phosphorylated and cleaved to produce C-terminal fragments [18,19,62,64–66]. The
variability in the distribution and morphology of ub-ir or TDP-43-ir inclusions in FTLD-U
has led to the development of a classification of FTLD-U into four pathological subtypes
based on the morphology of the inclusion, its location within the cell, and the density and
distribution in the frontal and/or temporal lobes (Fig. 3) [3,62,67]. These different
pathological subtypes correlate, although imperfectly, with the molecular genetic and
clinical phenotype [68]. Alternative schemes have been proposed and additional studies are
required to determine their validity [62,68]. The discovery of TDP-43 as a novel misfolded
protein links a spectrum of diseases including FTLD, ALS, Alzheimer’s disease [69], Lewy
body disease [70] and Guam parkinsonism-dementia complex [71] by a common molecular
pathology called TDP-43 proteinopathy. The identification of this protein as a major or
minor component of the pathological inclusions of a spectrum of neurodegenerative disease
provides new clues into the pathogenesis of protein aggregation, and it presents new targets
for therapeutic intervention where none exists.
TDP-43 proteinopathy is also a signature feature of other rare familial diseases. Inclusion
body myopathy associated with Paget’s disease of bone and frontotemporal dementia
(IBMPFD) is a rare autosomal dominant disorder caused by a mutation in the VCP gene
[32,33,62]. IBMPFD is a distinct subtype of FTLD-U with numerous DNs and NIIs [33,62].
As with cases of FTLD-U with GRN mutation, the ubiquitinated inclusions are not primarily
composed of the mutated protein VCP, but rather TDP-43 [62]. Recently, a new genetic
locus on chromosome 9p for familial ALS with or without FTLD has been described
[34,35]. Neuropathology of some of these families reveals the characteristic lesions of
FTLD-U: ub-ir and TDP-43-ir NCIs, DNs, and NIIs indicating that another gene locus on
chromosome 9 may precipitate TDP-43 proteinopathy [3].
Liscic et al. Page 4
Eur J Neurol. Author manuscript; available in PMC 2010 January 4.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
FTLD linked to chromosome 3 is an FTLD-U with mutation in the CHMP2B gene and is an
exception to the FTLD-U entities described above [31]. Early reports of this kindred
described the neuropathology as ‘dementia lacking distinctive histopathology’ (DLDH).
Subsequently, improved IHC methods revealed ub-ir, but TDP-43-negative, granular NCIs
in frontal neo-cortex and hippocampus [62,72]. Thus, rarely, FTLD-U may not be a TDP-43
proteinopathy. In the light of these recent immunohistochemical, biochemical, and genetic
advances, the Consortium for Frontotemporal Lobar Degeneration has developed an
algorithm to facilitate diagnosis and revised neuropathological diagnostic criteria (Table 1)
[3]. These criteria will be of value to the practicing clinician and provide a foundation for
clinical, clinico-pathological, and mechanistic studies and in vivo models of the pathogenesis
of FTLD.
Clinical evaluation
In addition to being the causative gene defect in autosomal dominantly inherited FTLD with
GRN mutation, a mutation in GRN may rarely occur in supposedly sporadic FTLD-U. Both
behavioral and aphasic forms of FTLD can be associated with these gene defects [45–
47,68], although the behavioral variant is the most common clinical phenotype associated
with FTLD with GRN mutation [46,47]. Magnetic resonance imaging and 18-fluoro-
deoxyglucose positron emission tomography helps discriminate AD from FTLD [73–77].
Patients with GRN mutations have predominant frontal, temporal and, to lesser extent,
parietal atrophy and hypometabolism with a right-sided predominance and this probably
relates to the predominance of behavioral symptoms [74]. However, language dysfunction in
patients with FTLD with GRN mutation show a left-sided predominance of atrophy on
imaging [45,78].
Several FTLD with GRN mutation families have been described: hereditary dysphasic
disinhibition dementia families 1 and 2 (HDDD1 and 2) [79,80] and other kindreds with a
similar clinical phenotype [37,75], UBC-17 [50,81], and aphasic families described by
Mesulam et al. [39] and Snowden et al. [82]. Interestingly, the HDDD1 and HDDD2
kindreds are characterized pathologically by FTLD-U and additional AD-type pathology,
which distinguishes them from other reported families with no or little coexisting
neurodegenerative disease [36,37]. Another family with the same GRN A9D mutation has
been reported in an individual with corticobasal syndrome [43] indicating, again, clinical
heterogeneity associated with the same mutation. Unlike most other FTLD-U with GRN
mutation families [47], the HDDD1 mutation carriers also had AD-type early-memory loss
which correlated with coexisting AD pathology [40]. The overlap between FTLD-U and AD
in familial cases is important as 23% of AD cases show FTLD-U type TDP-43 pathology
[67].
Recent discoveries in the molecular genetics, biochemistry, and neuropathology of FTLD
have transformed our understanding of this hitherto enigmatic group of diseases. TDP-43
has been identified as the major pathological protein of a spectrum of diseases which
includes FTLD-U, FTLD-U with ALS, and sporadic ALS, and a minor component of the
inclusions of a widening spectrum of other diseases. The physiological function of TDP-43
in the brain is currently unknown; however, it is normally localized to the nucleus of
neurons and some glial cells [19,64,65]. Although the specific role of TDP-43 in
neurodegeneration remains speculative, some studies indicate that this protein is directly
involved in the pathogenesis of SALS, ALS with dementia and FTLD-U [18,19,60,61]. The
most common FTLD is FTLD-U and accounts for more than one half of all FTLD entities in
most studies [51,52]. Ubiquitin and TDP-43 immunohistochemistry may be used to
distinguish four subtypes of FTLD-U which correlate with genotype and neuropathological
phenotype [62,67], but mutation analysis will be required to determine the genetic cause in
Liscic et al. Page 5
Eur J Neurol. Author manuscript; available in PMC 2010 January 4.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
familial cases. Additional studies of larger groups of cases are required to determine the
utility of the FTLD-U subtypes that have been described. The identification of different
mutations in the GRN gene in HDDD kindreds [38,40] links these families to other FTLD-U
families in which GRN mutations have been identified [36,37,43,50,68], but the HDDD
kindreds appear to be distinctive both clinically and pathologically. Further studies are
required to describe more accurately the clinical and neuropathological heterogeneity
associated with different gene defects that result in FTLD-U.
The identification of the causative gene defect in a patient with FTLD and/or ALS may
provide some insight into clinical and neuropathological variation that may be present.
Clinical studies of ALS, FTLD and FTLD with ALS have found significant clinical overlap,
indicating a spectrum of phenotypes, suggesting that they represent different manifestations
of the same neurodegenerative disorder [17,83]. The clinical overlap between ALS and
FTLD has also been illustrated in a study of 36 FTLD patients, five of whom met clinical
and electrophysiological criteria for definite ALS and an additional one-third met criteria for
possible ALS [15]. In contrast, in 100 consecutive ALS patients, frontal executive deficits
were present in half of ALS patients, many of whom met research criteria for FTLD [16]. In
patients with bulbar onset ALS, however, the incidence of FTLD has been reported as high
as 48% [14]. The presence of TDP-43 proteinopathy in cases with sporadic ALS, ALS with
FTLD, and FTLD links clinical phenotypes by a common molecular pathology. The absence
of pathological TDP-43, however, in ALS with SOD1 mutation may partially explain why
therapeutic strategies, shown to be effective in SOD1 mouse models, have not been effective
in clinical trials of patients with sporadic ALS [83–85].
Conclusion
FTLD and ALS are clinically, genetically, and neuropathologically heterogeneous. FTLD
can no longer be considered a rare group of disorders as it accounts for 5–10% of dementia
cases in most dementia centers. Although ALS and FTLD phenotypes may be distinguished
in the clinic, it is not usually possible to determine which neuropathological entity is
responsible for the clinical presentation. In familial cases, molecular genetics may identify
the causative gene but, at present, there is no effective treatment. Two distinct FTLD entities
have been linked to chromosome 17: FTLD-U with GRN mutation and FTLD with
microtubule-associated protein tau (MAPT) mutation. TDP-43 is the major pathological
protein of the motor neuron inclusions found in FTLD-U, FTLD-U with ALS, SALS, FALS,
but not FALS with SOD1 mutation, and G-PDC, and is a minor component of the inclusions
of other disorders including the neurofibrillary tangles of AD, Lewy bodies of Parkinson’s
disease and dementia with Lewy bodies, and the ubiquitinated inclusions of some cases of
hippocampal sclerosis. TDP-43 proteinopathies are distinct from most other
neurodegenerative disorders in which protein misfolding leads to brain amyloidosis, as
pathologic TDP-43 forms neuronal and glial inclusions lacking the features of brain amyloid
deposits [62]. Both HDDD kindreds are examples of FTLD-U with GRN mutation which
links them to other FTLD-U with GRN mutation families. However, the HDDD kindreds
appear to have coexisting memory deficits and AD-type pathology and as many as 20% AD
cases have pathology similar to FTLD-U indicating that these coexisting diseases may be
more frequent than previously realized. Also, TDP-43 proteinopathy may be a component of
a growing number of inclusions of different neurodegenerative diseases and further studies
are needed to determine the prevalence of this proteinopathy. Additional clinicopathological
correlations are required to discern the relationship between genotype, neuropathology, and
clinical phenotype more closely.
A rare form of familial FTLD-U is caused by mutations in the VCP gene, while other cases
of FTLD with ALS are linked to chromosome 9p. The majority of inclusions of sporadic and
Liscic et al. Page 6
Eur J Neurol. Author manuscript; available in PMC 2010 January 4.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
familial FTLD-U with GRN and VCP mutations contain TDP-43 protein, but FTLD-U with
CHMP2B mutation appears to be an exception with ub-ir positive and TDP-43-negative
inclusions. Also, familial cases of ALS with SOD1 mutations have ub-ir, but TDP-43-
negative inclusions. These data indicate that most, but not all, familial cases of FTLD or
ALS are TDP-43 proteinopathies. Thus, FTLD-U and SALS represent two ends of a
spectrum of disorders that are united by a common pathogenic mechanism TDP-43
proteinopathy which is reinforced by recent discoveries of pathogenic mutations in TARDBP
in some FALS. These dramatic new insights into the molecular genetics and neuropathology
of FTLD-U and ALS have led to a revised nosology of FTLD. These advances will
contribute toward an accurate diagnosis, foster clinicopathological studies, and facilitate the
quest for biomarkers and rational therapeutics.
Acknowledgments
We thank the clinical, genetic, pathology, and technical staff of the collaborating centers for making information
and tissue samples available for this study and we thank the families of patients whose generosity made this
research possible. This work was supported by NIH (National Institute on Aging) grants (P01-AG03991, P50-
AG05681, U01-AG16976), Fulbright Grant 68428174 (RML) and CAPES.
References
1. The Lund and Manchester Groups. Clinical and neuropathological criteria for frontotemporal
dementia. Journal of Neurology, Neurosurgery and Psychiatry. 1994; 57:416–418.
2. Neary D, Snowden JS, Gustafson L, et al. Frontotemporal lobar degeneration: a consensus on
clinical diagnostic criteria. Neurology. 1998; 51:1546–1554. [PubMed: 9855500]
3. Cairns NJ, Bigio EH, Mackenzie IR, et al. Neuropathologic diagnostic and nosologic criteria for
frontotemporal lobar degeneration: consensus of the Consortium for Frontotemporal Lobar
Degeneration. Acta Neuropathologica. 2007; 114:5–22. [PubMed: 17579875]
4. McKhann GM, Albert MS, Grossman M, Miller B, Dickson D, Trojanowski JQ. Clinical and
pathological diagnosis of frontotemporal dementia: report of the Work Group on Frontotemporal
Dementia and Pick’s Disease. Archives of Neurology. 2001; 58:1803–1809. [PubMed: 11708987]
5. Liscic RM, Storandt M, Cairns NJ, Morris JC. Clinical and psychometric distinction of
frontotemporal and Alzheimer dementias. Archives of Neurology. 2007; 64:535–540. [PubMed:
17420315]
6. Graham A, Davies R, Xuereb J, et al. Pathologically proven frontotemporal dementia presenting
with severe amnesia. Brain. 2005; 128:597–605. [PubMed: 15634737]
7. Neary D, Snowden JS, Mann DM. Classification and description of frontotemporal dementias.
Annals of the New York Academy of Sciences. 2000; 920:46–51. [PubMed: 11193176]
8. Ratnavalli E, Brayne C, Dawson K, Hodges JR. The prevalence of frontotemporal dementia.
Neurology. 2002; 58:1615–1621. [PubMed: 12058088]
9. Strong MJ, Kesavapany S, Pant HC. The pathobiology of ALS: a proteinopathy? Journal of
Neuropathology and Experimental Neurology. 2005; 64:649–664. [PubMed: 16106213]
10. Brooks BR. El Escorial World Federation of Neurology criteria for the diagnosis of amyotrophic
lateral sclerosis. Journal of the Neurological Sciences. 1994; 124(Suppl):96–107. [PubMed:
7807156]
11. Mitsuyama Y. Presenile dementia with motor neuron disease in Japan: clinicopathological review
of 26 cases. Journal of Neurology, Neurosurgery and Psychiatry. 1984; 47:953–959.
12. Neary D, Snowden JS, Mann DM, Northen B, Goulding PJ, Macdermott N. Frontal lobe dementia
and motor neuron disease. Journal of Neurology, Neurosurgery and Psychiatry. 1990; 53:23–32.
13. Massman PJ, Sims J, Cooke N, Haverkamp LJ, Appel V, Appel SH. Prevalence and correlates of
neuropsychological deficits in amyotrophic lateral sclerosis. Journal of Neurology, Neurosurgery
and Psychiatry. 1996; 61:450–455.
Liscic et al. Page 7
Eur J Neurol. Author manuscript; available in PMC 2010 January 4.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
14. Portet F, Cadilhac C, Touchon J, Camu W. Cognitive impairment in motor neuron disease with
bulbar onset. Amyotrophic Lateral Sclerosis and other Motor Neuron Disorders. 2001; 2:23–29.
[PubMed: 11465929]
15. Lomen-Hoerth C, Anderson T, Miller B. The overlap of amyotrophic lateral sclerosis and
frontotemporal dementia. Neurology. 2002; 59:1077–1079. [PubMed: 12370467]
16. Lomen-Hoerth C, Murphy J, Langmore S, Kramer JH, Olney RK, Miller B. Are amyotrophic
lateral sclerosis patients cognitively normal? Neurology. 2003; 60:1094–1097. [PubMed:
12682312]
17. Mackenzie IR, Feldman HH. Ubiquitin immunohistochemistry suggests classic motor neuron
disease, motor neuron disease with dementia, and frontotemporal dementia of the motor neuron
disease type represent a clinicopathologic spectrum. Journal of Neuropathology and Experimental
Neurology. 2005; 64:730–739. [PubMed: 16106222]
18. Arai T, Hasegawa M, Akiyama H, et al. TDP-43 is a component of ubiquitin-positive tau-negative
inclusions in frontotemporal lobar degeneration and amyotrophic lateral sclerosis. Biochemical
and Biophysical Research Communications. 2006; 351:602–611. [PubMed: 17084815]
19. Neumann M, Sampathu DM, Kwong LK, et al. Ubiquitinated TDP-43 in frontotemporal lobar
degeneration and amyotrophic lateral sclerosis. Science. 2006; 314:130–133. [PubMed: 17023659]
20. Caselli RJ, Windebank AJ, Petersen RC, et al. Rapidly progressive aphasic dementia and motor
neuron disease. Annals of Neurology. 1993; 33:200–207. [PubMed: 8257465]
21. Tsuchiya K, Ozawa E, Fukushima J, et al. Rapidly progressive aphasia and motor neuron disease: a
clinical, radiological, and pathological study of an autopsy case with circumscribed lobar atrophy.
Acta Neuropathologica. 2000; 99:81–87. [PubMed: 10651032]
22. Tsuchiya K, Ikeda K, Haga C, et al. Atypical amyotrophic lateral sclerosis with dementia
mimicking frontal Pick’s disease: a report of an autopsy case with a clinical course of 15 years.
Acta Neuropathologica. 2001; 101:625–630. [PubMed: 11515792]
23. Hodges JR, Davies R, Xuereb J, Kril J, Halliday G. Survival in frontotemporal dementia.
Neurology. 2003; 61:349–354. [PubMed: 12913196]
24. Josephs KA, Knopman DS, Whitwell JL, et al. Survival in two variants of tau-negative
frontotemporal lobar degeneration: FTLD-U vs FTLD-MND. Neurology. 2005; 65:645–647.
[PubMed: 16116138]
25. Gunnarsson LG, Dahlbom K, Strandman E. Motor neuron disease and dementia reported among 13
members of a single family. Acta Neurologica Scandinavica. 1991; 84:429–433. [PubMed:
1776392]
26. Chow TW, Miller BL, Hayashi VN, Geschwind DH. Inheritance of frontotemporal dementia.
Archives of Neurology. 1999; 56:817–822. [PubMed: 10404983]
27. Bird T, Knopman D, van Swieten J, et al. Epidemiology and genetics of frontotemporal dementia/
Pick’s disease. Annals of Neurology. 2003; 54(Suppl 5):S29–S31. [PubMed: 12833366]
28. Polvikoski TM, Murray A, Harper PS, Neal JW. Familial motor neurone disease with dementia:
phenotypic variation and cerebellar pathology. Journal of Neurology, Neurosurgery and
Psychiatry. 2003; 74:1516–1520.
29. Rosso SM, Donker KL, Baks T, et al. Frontotemporal dementia in The Netherlands: patient
characteristics and prevalence estimates from a population-based study. Brain. 2003; 126:2016–
2022. [PubMed: 12876142]
30. Martinaud O, Laquerriere A, Guyant-Marechal L, et al. Frontotemporal dementia, motor neuron
disease and tauopathy: clinical and neuropathological study in a family. Acta Neuropathologica.
2005; 110:84–92. [PubMed: 15965697]
31. Skibinski G, Parkinson NJ, Brown JM, et al. Mutations in the endosomal ESCRTIII-complex
subunit CHMP2B in frontotemporal dementia. Nature Genetics. 2005; 37:806–808. [PubMed:
16041373]
32. Watts GD, Wymer J, Kovach MJ, et al. Inclusion body myopathy associated with Paget disease of
bone and frontotemporal dementia is caused by mutant valosin-containing protein. Nature
Genetics. 2004; 36:377–381. [PubMed: 15034582]
Liscic et al. Page 8
Eur J Neurol. Author manuscript; available in PMC 2010 January 4.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
33. Forman MS, Mackenzie IR, Cairns NJ, et al. Novel ubiquitin neuropathology in frontotemporal
dementia with valosin-containing protein gene mutations. Journal of Neuropathology and
Experimental Neurology. 2006; 65:571–581. [PubMed: 16783167]
34. Morita M, Al Chalabi A, Andersen PM, et al. A locus on chromosome 9p confers susceptibility to
ALS and frontotemporal dementia. Neurology. 2006; 66:839–844. [PubMed: 16421333]
35. Vance C, Al Chalabi A, Ruddy D, et al. Familial amyotrophic lateral sclerosis with frontotemporal
dementia is linked to a locus on chromosome 9p13.2–21.3. Brain. 2006; 129:868–876. [PubMed:
16495328]
36. Baker M, Mackenzie IR, Pickering-Brown SM, et al. Mutations in progranulin cause tau-negative
frontotemporal dementia linked to chromosome 17. Nature. 2006; 442:916–919. [PubMed:
16862116]
37. Cruts M, Gijselinck I, van der Zee J, et al. Null mutations in progranulin cause ubiquitin-positive
frontotemporal dementia linked to chromosome 17q21. Nature. 2006; 442:920–924. [PubMed:
16862115]
38. Mukherjee O, Pastor P, Cairns NJ, et al. HDDD2 is a familial frontotemporal lobar degeneration
with ubiquitin-positive, tau-negative inclusions caused by a missense mutation in the signal
peptide of progranulin. Annals of Neurology. 2006; 60:314–322. [PubMed: 16983685]
39. Mesulam M, Johnson N, Krefft TA, et al. Progranulin mutations in primary progressive aphasia-
the PPA1 and PPA3 families. Archives of Neurology. 2007; 64:314–322.
40. Behrens MI, Mukherjee O, Tu PH, et al. Neuropathologic heterogeneity in HDDD1: a familial
frontotemporal lobar degeneration with ubiquitin-positive inclusions and progranulin mutation.
Alzheimer Disease and Associated Disorders. 2007; 21:1–7. [PubMed: 17334266]
41. Benussi L, Binetti G, Sina E, et al. A novel deletion in progranulin gene is associated with
FTDP-17 and CBD. Neurobiology of Aging. 2008; 29:427–435. [PubMed: 17157414]
42. Masselis M, Momeni P, Meschino W, et al. Novel splicing mutation in the progranulin gene
causing familial corticobasal syndrome. Brain. 2006; 129:3115–3123. [PubMed: 17030534]
43. Spina S, Murrell J, Huey E, et al. Corticobasal syndrome associated with the A9D progranulin
mutation. Journal of Neuropathology and Experimental Neurology. 2007; 66:892–900. [PubMed:
17917583]
44. Gass J, Cannon A, Mackenzie IR, et al. Mutations in progranulin are a major cause of ubiquitin-
positive frontotemporal lobar degeneration. Human Molecular Genetics. 2006; 15:2988–3001.
[PubMed: 16950801]
45. Huey ED, Grafman J, Wassermann EM, et al. Characteristics of frontotemporal dementia patients
with a progranulin mutation. Annals of Neurology. 2006; 60:374–380. [PubMed: 16983677]
46. Josephs KA, Ahmed Z, Katsuse O, et al. Neuropathologic features of frontotemporal lobar
degeneration with ubiquitin-positive inclusions with progranulin gene (PGRN) mutations. Journal
of Neuropathology and Experimental Neurology. 2006; 66:142–151. [PubMed: 17278999]
47. Rademakers R, Baker M, Gass J, et al. Phenotypic variability associated with progranulin
haploinsufficiency in patients with the common 1477C–>T (Arg493X) mutation: an international
initiative. Lancet Neurology. 2007; 6:857–868. [PubMed: 17826340]
48. Spina S, Murrell JR, Huey ED, et al. Clinicopathologic features of frontotemporal dementia with
progranulin sequence variation. Neurology. 2007; 68:820–827. [PubMed: 17202431]
49. Mukherjee O, Wang J, Gitcho M, et al. A Molecular characterization of novel progranulin (GRN)
mutations in frontotemporal dementia. Human Mutation. 2008; 29:512–521. [PubMed: 18183624]
50. Mackenzie IR, Baker M, Pickering-Brown S, et al. The neuropathology of frontotemporal lobar
degeneration caused by mutations in the progranulin gene. Brain. 2006; 129:3081–3090. [PubMed:
17071926]
51. Ikeda M, Ishikawa T, Tanabe H. Epidemiology of frontotemporal lobar degeneration. Dementia
and Geriatric Cognitive Disorders. 2004; 17:265–268. [PubMed: 15178933]
52. Rademakers R, Cruts M, van Broeckhoven C. Epidemiology of frontotemporal dementia and
related tauopathies. Human Mutation. 2004; 24:277–295. [PubMed: 15365985]
53. Josephs KA, Holton JL, Rossor MN, et al. Frontotemporal lobar degeneration and ubiquitin
immunohistochemistry. Neuropathology and Applied Neurobiology. 2004; 30:369–373. [PubMed:
15305982]
Liscic et al. Page 9
Eur J Neurol. Author manuscript; available in PMC 2010 January 4.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
54. Gitcho MA, Baloh RH, Chakraverty S, et al. TDP-43 A315T mutation in familial motor neuron
disease. Annals of Neurology. 2008; 63:535–538. [PubMed: 18288693]
55. Sreedharan J, Blair IP, Tripathi VB, et al. TDP-43 mutations in familial and sporadic amyotrophic
lateral sclerosis. Science. 2008; 319:1668–1672. [PubMed: 18309045]
56. Kabashi E, Valdmanis PN, Dion P, et al. TARDBP mutations in individuals with sporadic and
familial amyotrophic lateral sclerosis. Nature Genetics. 2008; 40:572–574. [PubMed: 18372902]
57. Van Deerlin VM, Leverenz JB, Bekris LM, et al. TARDBP mutations in amyotrophic lateral
sclerosis with TDP-43 neuropathology: a genetic and histopathological analysis. Lancet
Neurology. 2008; 5:409–416. [PubMed: 18396105]
58. Lipton AM, White CL III, Bigio EH. Frontotemporal lobar degeneration with motor neuron
disease-type inclusions predominates in 76 cases of frontotemporal degeneration. Acta
Neuropathologica. 2004; 108:379–385. [PubMed: 15351890]
59. Leigh PN, Whitwell H, Garofalo O, et al. Ubiquitin-immunoreactive intraneuronal inclusions in
amyotrophic lateral sclerosis. Morphology, distribution, and specificity. Brain. 1991; 114:775–
788. [PubMed: 1646064]
60. Mackenzie IR, Bigio EH, Ince PG, et al. Pathological TDP-43 distinguishes sporadic amyotrophic
lateral sclerosis from amyotrophic lateral sclerosis with SOD1 mutations. Annals of Neurology.
2007; 61:427–434. [PubMed: 17469116]
61. Tan CF, Eguchi H, Tagawa A, et al. TDP-43 immunoreactivity in neuronal inclusions in familial
amyotrophic lateral sclerosis with or without SOD1 gene mutations. Acta Neuropathologica. 2007;
113:535–542. [PubMed: 17333220]
62. Cairns NJ, Neumann M, Bigio EH, et al. TDP-43 in familial and sporadic frontotemporal lobar
degeneration with ubiquitin inclusions. American Journal of Pathology. 2007; 171:227–240.
[PubMed: 17591968]
63. Woulfe J, Kertesz A, Munoz DG. Frontotemporal dementia with ubiquitinated cytoplasmic and
intranuclear inclusions. Acta Neuropathologica. 2001; 102:94–102. [PubMed: 11547957]
64. Neumann M, Kwong LK, Truax AC, et al. TDP-43-positive white matter pathology in
frontotemporal lobar degeneration with ubiquitin-positive inclusions. Journal of Neuropathology
and Experimental Neurology. 2007; 66:177–183. [PubMed: 17356379]
65. Neumann M, Kwong LK, Sampathu DM, Trojanowski JQ, Lee VM-Y. TDP-43 proteinopathy in
frontotemporal lobar degeneration and amyotrophic lateral sclerosis. Archives of Neurology. 2007;
64:1388–1394. [PubMed: 17923623]
66. Neumann M, Mackenzie IR, Cairns NJ, et al. TDP-43 in the ubiquitin pathology of frontotemporal
dementia with VCP gene mutations. Journal of Neuropathology and Experimental Neurology.
2007; 66:152–157. [PubMed: 17279000]
67. Sampathu DM, Neumann M, Kwong LK, et al. Pathological heterogeneity of frontotemporal lobar
degeneration with ubiquitin-positive inclusions delineated by ubiquitin immunohistochemistry and
novel monoclonal antibodies. American Journal of Pathology. 2006; 169:1343–1352. [PubMed:
17003490]
68. Mackenzie IRA, Rademakers R. The molecular genetics and neuropathology of frontotemporal
lobar degeneration: recent developments. Neurogenetics. 2007; 8:237–248. [PubMed: 17805587]
69. Amador-Ortiz C, Lin W-L, Ahmed Z, et al. TDP-43 immunoreactivity in hippocampal sclerosis
and Alzheimer’s disease. Annals of Neurology. 2007; 61:435–445. [PubMed: 17469117]
70. Nakashima-Yasuda H, Uryu K, Robinson J, et al. Co-morbidity of TDP-43 proteinopathy in Lewy
body related diseases. Acta Neuropathologica (Berlin). 2007; 114:221–230. [PubMed: 17653732]
71. Hasegawa M, Arai T, Akiyama H, et al. TDP-43 is deposited in the Guam parkinsonism-dementia
complex brains. Brain. 2007; 130:1386–1394. [PubMed: 17439983]
72. Holm IE. Ubiquitin-positive inclusions in frontotemporal dementia linked to chromosome 3
(FTD-3). Brain Pathology. 2006; 16(Suppl 1):S43.
73. Davion S, Johnson N, Weintraub S, et al. Clinicopathological correlation in PGRN mutations.
Neurology. 2007; 69:1113–1121. [PubMed: 17522386]
74. Joeng Y, Cho SS, Park JM, et al. 18F-FDG PET findings in frontotemporal dementia: an SPM
analysis of 29 patients. Journal of Nuclear Medicine. 2005; 46:233–239. [PubMed: 15695781]
Liscic et al. Page 10
Eur J Neurol. Author manuscript; available in PMC 2010 January 4.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
75. Rabinovici GD, Furst AJ, O’Neil JP, et al. 11C-PIB PET imaging in Alzheimer disease and
frontotemporal lobar degeneration. Neurology. 2007; 68:1205–1212. [PubMed: 17420404]
76. Whitwell JL, Clifford RJ Jr, Baker M, et al. Voxel-based morphometry in frontotemporal lobar
degeneration with ubiquitin-positive inclusions with and without progranulin mutations. Archives
of Neurology. 2007; 64:371–376. [PubMed: 17353379]
77. Rosen HJ, Allison SC, Schauer GF, et al. Neuroanatomical correlates of behavioral disorders in
dementia. Brain. 2005; 128:2612–2625. [PubMed: 16195246]
78. van der Zee J, Rademakers R, Engelborghs S, et al. A Belgian ancestral haplotype harbours a
highly prevalent mutation for 17q21-linked tau-negative FTLD. Brain. 2006; 129:841–852.
[PubMed: 16495329]
79. Morris JC, Cole M, Banker BQ, Wright D. Hereditary dysphasic dementia and the Pick-Alzheimer
spectrum. Annals of Neurology. 1984; 16:455–466. [PubMed: 6497355]
80. Lendon CL, Lynch T, Norton J, et al. Hereditary dysphasic disinhibition dementia: a
frontotemporal dementia linked to 17q21-22. Neurology. 1998; 50:1546–1555. [PubMed:
9633693]
81. Mackenzie IR, Baker M, West G, et al. A family with tau-negative frontotemporal dementia and
neuronal intranuclear inclusions linked to chromosome 17. Brain. 2006; 129:853–867. [PubMed:
16401619]
82. Snowden JS, Pickering-Brown SM, Mackenzie IR, et al. Progranulin gene mutations associated
with frontotemporal dementia and progressive non-fluent aphasia. Brain. 2006; 129:3091–3102.
[PubMed: 17003069]
83. Strong MJ, Lomen-Hoerth C, Caselli RJ, Bigio EH, Yang W. Cognitive impairment,
frontotemporal dementia, and the motor neuron diseases. Annals of Neurology. 2003; 54(Suppl
5):S20–S23. [PubMed: 12833364]
84. Ludolph AC, Sperfeld AD. Preclinical trials: an update on translational research in ALS. Neuro-
degenerative Diseases. 2005; 2:215–219. [PubMed: 16909028]
85. DiBernardo AB, Cudkowicz ME. Translating preclinical insights into effective human trials in
ALS. Biochimica et Biophysica Acta. 2006; 1762:1139–1149. [PubMed: 16713196]
Liscic et al. Page 11
Eur J Neurol. Author manuscript; available in PMC 2010 January 4.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
Figure 1.
Mutations in the TAR DNA-binding protein 43 (TDP-43) encoded by the TARDBP gene in
familial (FALS) and sporadic amyotrophic lateral sclerosis (SALS). Schematic diagram of
functional domains and mutations in the coding region are indicated using the amino acid
numbering of the 414 amino acid protein. RRM1, RNA recognition motif 1; RRM2, RNA
recognition motif 2; NL, nuclear localization signal; NE, nuclear export signal; hnRNP,
heterogeneous nuclear ribonucleoprotein interaction domain. TARDBP mutations:
autosomal-dominant FALS (black), SALS (green).
Liscic et al. Page 12
Eur J Neurol. Author manuscript; available in PMC 2010 January 4.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
Figure 2.
TDP-43 proteinopathy in FTLD-U. Adjacent sections of superficial frontal neocortex
showing neuronal cytoplasmic inclusions (NCIs), dystrophic neurites (DNs), and isolated
neuronal intranuclear inclusions (NIIs), stained for both ubiquitin (a) and TDP-43 (b). NCIs
in the dentate granule cells stain for ubiquitin (c) and TDP-43 (d). Neuronal and glial
inclusions include: NCI (e), round and lentiform NIIs (f, g); skein-like (h) and compact
round (i) NCIs in lower motor neurons; and a glial cytoplasmic inclusion (GCI) (j). (a, c)
ubiquitin immunohistochemisty; (b, d, e–j) TDP-43 immunohistochemistry. Bars 10 μm (a–
d); 5 μm (e–j). Source: Cairns et al. [60] with permission from the American Society for
Investigative Pathology.
Liscic et al. Page 13
Eur J Neurol. Author manuscript; available in PMC 2010 January 4.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
Figure 3.
FTLD-U subtypes 1–4. (a) Type 1 is characterized by long dystrophic neurites (DNs) in
laminae II/III with relatively few neuronal cytoplasmic inclusions (NCIs) and no neuronal
intranuclear inclusion (NII). (b) Type 2 has numerous NCIs, relatively few DNs, and no NII.
(c) Type 3 has numerous NCIs and DNs and an occasional NII. (d) Type 4 pathology is
characterized by numerous NIIs and DNs but few NCIs. TDP-43 immunohistochemistry.
Bar 10 μm (a–d). Source: Cairns et al. [60] with permission from the American Society for
Investigative Pathology.
Liscic et al. Page 14
Eur J Neurol. Author manuscript; available in PMC 2010 January 4.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
Liscic et al. Page 15
Table 1
Neuropathologic diagnostic criteria for FTLD
1Tauopathy (with associated neuron loss and gliosis) and insoluble tau with a predominance of 3R tau, the most likely diagnoses are:
FTLD with Pick bodies
FTLD with MAPT mutation
2Tauopathy (with associated neuron loss and gliosis) and insoluble tau with a predominance of 4R tau, the most likely diagnoses are:
Corticobasal degeneration
Progressive supranuclear palsy
Argyrophilic grain disease
Sporadic multiple system tauopathy with dementia
FTLD with MAPT mutation
3Tauopathy (with associated neuron loss and gliosis) and insoluble tau with a predominance of 3R and 4R tau, the most likely
diagnoses are:
Neurofibrillary tangle dementia
FTLD with MAPT mutation
4Frontotemporal neuronal loss and gliosis without tau- or ubiquitin/p62- positive inclusions, the most likely diagnoses is:
FTLD (also known as dementia lacking distinctive histologic features)
5TDP-43 proteinopathy with associated neuronal loss and ubiquitin-positive/p62-positive, tau-negative inclusions, with MND or
without MND but with MND-type inclusions, the most likely diagnoses are:
FTLD-U with MND (FTLD-U types 1–3)
FTLD-U but without MND (FTLD-U types 1–3)
FTLD-U with GRN mutation (FTLD-U type 3)
FTLD-U with VCP mutation (FTLD-U type 4)
FTLD-U linked to chromosome 9p (FTLD-U type 2)
Other as yet unidentified TDP-43 proteinopathies
6Frontotemporal neuronal loss and gliosis with ubiquitin-positive/p62-positive, TDP-43- and tau-negative inclusions, the most likely
diagnoses are:
FTLD-U with CHMP2B mutation
Basophilic inclusion body disease (BIBD)
Other as yet unidentified FTLD-U, non-TDP-43 proteinopathies
7Frontotemporal neuronal loss and gliosis with ubiquitin/p62 and α-internexin-positive inclusions, the most likely diagnosis is:
Neuronal intermediate filament inclusion disease (NIFID)
CHMP2B, charged multivesicular body protein 2B gene; FTLD, frontotemporal lobar degeneration; FTLD-U, FTLD with ubiquitin-positive, tau-,
α-synuclein-, TDP-43-, and neuronal intermediate filament protein-negative inclusions; MAPT, microtubule-associated protein tau gene; MND,
motor neuron disease; neurofibrillary tangle dementia, also called tangle predominant form of senile dementia; GRN, granulin gene; TDP-43, TAR
DNA-binding protein 43; VCP, valosin-containing protein gene. Reprinted from Ref. [3] with permission of the Editor.
Eur J Neurol. Author manuscript; available in PMC 2010 January 4.
... Only participants with brain TDP-43 who did not have ALS or FTLD-TDP or the aforementioned TDP-43 related pathologies were then classified as LATE-NC. In the data used, ALS-TDP often co-occurred with FTLD-TDP, and so we opted to group ALS-TDP and FTLD-TDP into a single entity of ALS/FTLD-TDP, which is consistent with the concept that these two pathologies exist on an ALS/FTLD spectrum [19,49]. In Fig. 1b we show diagrams for how the TDP-43 categories were assigned to participants with all TDP-43 related information available. ...
... In this study we found a high overlap of ALS-TDP and FTLD-TDP, which concurs with the idea that ALS-TDP and FTLD-TDP are often on the same spectrum [19,28,49]. Participants with both ALS-TDP and FTLD-TDP tended to die at younger ages than those with either pathology alone. ...
Article
Full-text available
TDP-43 proteinopathy is a salient neuropathologic feature in a subset of frontotemporal lobar degeneration (FTLD-TDP), in amyotrophic lateral sclerosis (ALS-TDP), and in limbic-predominant age-related TDP-43 encephalopathy neuropathologic change (LATE-NC), and is associated with hippocampal sclerosis of aging (HS-A). We examined TDP-43-related pathology data in the National Alzheimer’s Coordinating Center (NACC) in two parts: (I) availability of assessments, and (II) associations with clinical diagnoses and other neuropathologies in those with all TDP-43 measures available. Part I: Of 4326 participants with neuropathology data collected using forms that included TDP-43 assessments, data availability was highest for HS-A (97%) and ALS (94%), followed by FTLD-TDP (83%). Regional TDP-43 pathologic assessment was available for 77% of participants, with hippocampus the most common region. Availability for the TDP-43-related measures increased over time, and was higher in centers with high proportions of participants with clinical FTLD. Part II: In 2142 participants with all TDP-43-related assessments available, 27% of participants had LATE-NC, whereas ALS-TDP or FTLD-TDP (ALS/FTLD-TDP) was present in 9% of participants, and 2% of participants had TDP-43 related to other pathologies (“Other TDP-43”). HS-A was present in 14% of participants, of whom 55% had LATE-NC, 20% ASL/FTLD-TDP, 3% Other TDP-43, and 23% no TDP-43. LATE-NC, ALS/FTLD-TDP, and Other TDP-43, were each associated with higher odds of dementia, HS-A, and hippocampal atrophy, compared to those without TDP-43 pathology. LATE-NC was associated with higher odds for Alzheimer’s disease (AD) clinical diagnosis, AD neuropathologic change (ADNC), Lewy bodies, arteriolosclerosis, and cortical atrophy. ALS/FTLD-TDP was associated with higher odds of clinical diagnoses of primary progressive aphasia and behavioral-variant frontotemporal dementia, and cortical/frontotemporal lobar atrophy. When using NACC data for TDP-43-related analyses, researchers should carefully consider the incomplete availability of the different regional TDP-43 assessments, the high frequency of participants with ALS/FTLD-TDP, and the presence of other forms of TDP-43 pathology. Supplementary Information The online version contains supplementary material available at 10.1007/s00401-024-02728-8.
... TDP-43 pathology is one of the most common pathologies observed in a broad spectrum of ALS patients, including familial (fALS), sporadic (sALS) and ALS/frontotemporal lobar degeneration (ALS/FTLD) [11][12][13][14][15][16][17], which is accompanied by progressive UMN loss and increased astrogliosis and microgliosis, especially in layer 5 of the motor cortex. Despite their importance, there is currently no treatment strategies to improve the health of UMNs, or to reduce astrogliosis and microgliosis in the motor cortex, -two important contributors to motor neuron degeneration with respect to TDP-43 pathology [18][19][20][21][22][23]. ...
... TDP-43 pathology is one of the most common proteinopathies, detected in the CNS of a broad spectrum of ALS patients, including sporadic [11], familial [12], and mostly ALS/FTLD [13][14][15][16][17]. TDP-43 pathology is defined by the presence of protein accumulations that contain phosphorylated TDP-43 and ubiquitinated proteins, mostly in the cytoplasm [73,74], and sometimes in the nucleus [75,76]. ...
Article
Full-text available
Recombinant adeno-associated virus (rAAV)-based gene therapies offer an immense opportunity for rare diseases, such as amyotrophic lateral sclerosis (ALS), which is defined by the loss of the upper and the lower motor neurons. Here, we describe generation, characterization, and utilization of a novel vector system, which enables expression of the active form of hepatocyte growth factor (HGF) under EF-1α promoter with bovine growth hormone (bGH) poly(A) sequence and is effective with intrathecal injections. HGF’s role in promoting motor neuron survival had been vastly reported. Therefore, we investigated whether intrathecal delivery of HGF would have an impact on one of the most common pathologies of ALS: the TDP-43 pathology. Increased astrogliosis, microgliosis and progressive upper motor neuron loss are important consequences of ALS in the motor cortex with TDP-43 pathology. We find that cortex can be modulated via intrathecal injection, and that expression of HGF reduces astrogliosis, microgliosis in the motor cortex, and help restore ongoing UMN degeneration. Our findings not only introduce a novel viral vector for the treatment of ALS, but also demonstrate modulation of motor cortex by intrathecal viral delivery, and that HGF treatment is effective in reducing astrogliosis and microgliosis in the motor cortex of ALS with TDP-43 pathology.
... The association between TBI and ALS/FTD is further exemplified by the shared neuropathology of TAR DNA-binding protein aggregation [37][38][39][40]. TDP-43 is a predominantly nuclear RNA binding protein involved in RNA transport and splicing, that becomes depleted from the nucleus and forms cytoplasmic aggregations in the overwhelming majority of ALS (~97%) and as many as 50% of all FTD cases [41][42][43]. Notably, widespread TDP-43 aggregation is also seen in the brain and spinal cord of patients who have suffered TBI and CTE [44][45][46]. While these observations indicate a potential link between these diverse neurodegenerative diseases, it is unclear whether . ...
Preprint
Full-text available
A hexanucleotide repeat expansion (HRE) in C9orf72 is the most common genetic cause of amyotrophic lateral sclerosis (ALS) and frontotemporal dementia (FTD). However, patients with the HRE exhibit a wide disparity in clinical presentation and age of symptom onset suggesting an interplay between genetic background and environmental stressors. Neurotrauma as a result of traumatic brain or spinal cord injury has been shown to increase the risk of ALS/FTD in epidemiological studies. Here, we combine patient-specific induced pluripotent stem cells (iPSCs) with a custom-built device to deliver biofidelic stretch trauma to C9orf72 patient and isogenic control motor neurons (MNs) in vitro . We find that mutant but not control MNs exhibit selective degeneration after a single incident of severe trauma, which can be partially rescued by pretreatment with a C9orf72 antisense oligonucleotide. A single incident of mild trauma does not cause degeneration but leads to cytoplasmic accumulation of TDP-43 in C9orf72 MNs. This mislocalization, which only occurs briefly in isogenic controls, is eventually restored in C9orf72 MNs after 6 days. Lastly, repeated mild trauma ablates the ability of patient MNs to recover. These findings highlight alterations in TDP-43 dynamics in C9orf72 ALS/FTD patient MNs following traumatic injury and demonstrate that neurotrauma compounds neuropathology in C9orf72 ALS/FTD. More broadly, our work establishes an in vitro platform that can be used to interrogate the mechanistic interactions between ALS/FTD and neurotrauma.
... In ALS, TDP-43 inclusions accumulate in the motoneurons of the primary motor cortex, corticospinal tracts, brainstem, spinal cord, glial cells, and also in the neurons of the hippocampus and frontotemporal cortex in a subset of patients [2,5,7]. In FTLD-U the NCIs are mainly present in the frontal and temporal cortices and in the hippocampus and their glial cells [2,8]. TDP-43 inclusions are also frequently found in the brains of Alzheimer's disease [9,10], Parkinson's disease [9,11], Huntington's disease [12], and other neurodegenerative conditions [13], although in this case they are thought to represent a secondary effect. ...
Article
Full-text available
Introduction: Several neurodegenerative conditions are associated with a common histopathology within neurons of the central nervous system, consisting of the deposition of cytoplasmic inclusions of TAR DNA-binding protein 43 (TDP-43). Such inclusions have variably been described as morphologically and molecularly ordered aggregates having amyloid properties, as filaments without the cross-β-structure and dye binding specific for amyloid, or as amorphous aggregates with no defined structure and fibrillar morphology. Aims and Methods: Here we have expressed human full-length TDP-43 in neuroblastoma x spinal cord 34 (NSC-34) cells to investigate the morphological, structural, and tinctorial properties of TDP-43 inclusions in situ. We have used last-generation amyloid diagnostic probes able to cross the cell membrane and detect amyloid in the cytoplasm and have adopted Raman and Fourier transform infrared microspectroscopies to study in situ the secondary structure of the TDP-43 protein in the inclusions. We have then used transmission electron microscopy to study the morphology of the TDP-43 inclusions. Results: The results show the absence of amyloid dye binding, the lack of an enrichment of cross-β structure in the inclusions, and of a fibrillar texture in the round inclusions. The aggregates formed in vitro from the purified protein under conditions in which it is initially native also lack all these characteristics, ruling out a clear amyloid-like signature. Conclusions: These findings indicate a low propensity of TDP-43 to form amyloid fibrils and even non-amyloid filaments, under conditions in which the protein is initially native and undergoes its typical nucleus-to-cell mislocalization. It cannot be excluded that filaments emerge on the long time scale from such inclusions, but the high propensity of the protein to form initially other types of inclusions appear to be an essential characteristic of TDP-43 proteinopathies. KEY MESSAGES Cytoplasmic inclusions of TDP-43 formed in NSC-34 cells do not stain with amyloid-diagnostic dyes, are not enriched with cross-β structure, and do not show a fibrillar morphology. TDP-43 assemblies formed in vitro from pure TDP-43 do not have any hallmarks of amyloid.
... Clinically, sporadic and familial types do not distinguish between each other, but in comparison with sporadic, the familial type of the disease occurs at an earlier age of life [35], approximately a decade before the sporadic type [31], lasts longer, and women and men get ill with the same frequency [35,36]. Up to 50% of all patients with ALS also express symptoms of frontotemporal dementia (FTD), which leads to the degradation of neurons in frontotemporal brain lobes [37]. Interestingly, it was discovered that some of the ALS-specific genes, UBQLN2 [21] and C9orf72 [28,29], are responsible for the combination of ALS-FTD [38]. ...
Article
Full-text available
Neurodegenerative diseases are one of the greatest medical burdens of the modern age, being mostly incurable and with limited prognostic and diagnostic tools. Amyotrophic lateral sclerosis (ALS) is a fatal, progressive neurodegenerative disease characterized by the loss of motoneurons, with a complex etiology, combining genetic, epigenetic, and environmental causes. The neuroprotective therapeutic approaches are very limited, while the diagnostics rely on clinical examination and the exclusion of other diseases. The recent advancement in the discovery of molecular pathways and gene mutations involved in ALS has deepened the understanding of the disease pathology and opened the possibility for new treatments and diagnostic procedures. Recently, 15 risk loci with distinct genetic architectures and neuron-specific biology were identified as linked to ALS through common and rare variant association analyses. Interestingly, the quantity of related proteins to these genes has been found to change during early postnatal development in mammalian spinal cord tissue (opossum Monodelphis domestica) at the particular time when neuroregeneration stops being possible. Here, we discuss the possibility that the ALS-related genes/proteins could be connected to neuroregeneration and development. Moreover, since the regulation of gene expression in developmental checkpoints is frequently regulated by non-coding RNAs, we propose that studying the changes in the composition and quantity of non-coding RNA molecules, both in ALS patients and in the developing central nervous (CNS) system of the opossum at the time when neuroregeneration ceases, could reveal potential biomarkers useful in ALS prognosis and diagnosis.
Article
Full-text available
INTRODUCTION TAR DNA‐binding protein 43 (TDP‐43) is a highly prevalent proteinopathy that is involved in neurodegenerative processes, including axonal damage. To date, no ante mortem biomarkers exist for TDP‐43, and few studies have directly assessed its impact on neuroimaging measures utilizing pathologic quantification. METHODS Ante mortem diffusion‐weighted images were obtained from community‐dwelling older adults. Regression models calculated the relationship between post mortem TDP‐43 burden and ante mortem fractional anisotropy (FA) within each voxel in connection with the hippocampus, controlling for coexisting Alzheimer's disease and demographics. RESULTS Results revealed a significant negative relationship (false discovery rate [FDR] corrected p < .05) between post mortem TDP‐43 and ante mortem FA in one cluster within the left medial temporal lobe connecting to the parahippocampal cortex, entorhinal cortex, and cingulate, aligning with the ventral subdivision of the cingulum. FA within this cluster was associated with cognition. DISCUSSION Greater TDP‐43 burden is associated with lower FA within the limbic system, which may contribute to impairment in learning and memory. Highlights Post mortem TDP‐43 pathological burden is associated with reduced ante mortem fractional anisotropy. Reduced FA located in the parahippocampal portion of the cingulum. FA in this area was associated with reduced episodic and semantic memory. FA in this area was associated with increased inward hippocampal surface deformation.
Article
Amyotrophic lateral sclerosis is a fatal degenerative neurological disease known as motor neuron disease. According to research, amyotrophic lateral sclerosis affects nerve cells related to movement in the brain and spinal cord, leading to the death of motor neurons, making the brain unable to control muscle movement, and resulting in a significant loss of motor nerve cells, leading to muscle atrophy. In the early stages of ALS, symptoms are mild, often go unnoticed, and can easily be confused with other diseases. Patients may only feel symptoms such as weakness, convulsions, and fatigue, gradually develop muscle atrophy all over the body, difficulty swallowing, and finally die of respiratory failure. TDP-43 is a DNA/RNA binding protein typically located in the nucleus regulating various steps of RNA metabolism. Meanwhile, TDP-43 is the most prominent pathological protein in the characteristics of ALS patients.[1] TDP-43 is depleted in the nucleus in nearly all ALS cases but accumulates in the cytoplasm.Deep learning is learning the internal laws and representation levels of sample data. The information obtained from these learning processes can significantly help interpret data such as text, images, and sounds. Its goal is to enable machines to analyze and learn like humans and recognize data. Additionally, super-resolution microscopy can be used to observe the conformation of proteins in biological samples, thereby gaining insight into the structure and assembly mechanism of TDP-43 aggregates.This project aims to use super-resolution images for machine learning to build a model for classifying ALS patients from non-ALS patients. At the same time, multiple neural network models are used for training and then compared to select the model with the highest accuracy.
Article
Distal hereditary neuropathies and neuro motor diseases are complex neurological phenotypes associated with pathogenic variants in a large number of genes, but in some the origin is unknown. Recently, rare pathogenic variants of the human VRK1 gene have been associated with these neurological phenotypes. All VRK1 pathogenic variants are recessive, and their clinical presentation occurs in either homozygous or compound heterozygous patients. The pathogenic VRK1 gene pathogenic variants are located in three clusters within the protein sequence. The main, and initial, shared clinical phenotype among VRK1 pathogenic variants is a distal progressive loss of motor and/or sensory function, which includes diseases such as spinal muscular atrophy, Charcot-Marie-Tooth, amyotrophic lateral sclerosis and hereditary spastic paraplegia. In most cases, symptoms start early in infancy, or in utero, and are slowly progressive. Additional neurological symptoms vary among non-related patients, probably because of their different VRK1 variants and their genetic background. The underlying common pathogenic mechanism, by its functional impairment, is a likely consequence of the roles that the VRK1 protein plays in the regulation on the stability and assembly of Cajal bodies, which affect RNA maturation and processing, neuronal migration of RNPs along axons, and DNA-damage responses. Alterations of these processes are associated with several neuro sensory or motor syndromes. The clinical heterogeneity of the neurological phenotypes associated with VRK1 is a likely consequence of the protein complexes in which VRK1 is integrated, which include several proteins known to be associated with Cajal bodies and DNA damage responses. Several hereditary distal neurological diseases are a consequence of pathogenic variants in genes that alter these cellular functions. We conclude that VRK1-related distal hereditary neuropathies and motor neuron diseases represent a novel subgroup of Cajal body related neurological syndromes.
Article
We investigated whether pharmacological increase of "M-type" (KCNQ, Kv7) K + channel currents by the M-channel opener, retigabine (RTG), acutely after repetitive traumatic brain injuries (rTBIs), prevents or reduces their long-term detrimental effects. rTBIs were studied using a blast shock air wave mouse model. Animals were monitored by video and electroencephalogram (EEG) records for nine months after the last injury to assess the occurrence of post-traumatic seizures (PTS), post-traumatic epilepsy (PTE), sleep-wake cycle architecture alterations, and the power of the EEG signals. We evaluated the development of long-term changes in the brain associated with various neurodegenerative diseases in mice by examining transactive response DNA-binding protein 43 (TDP-43) expression and nerve fiber damage ~ 2 years after the rTBIs. We observed acute RTG treatment to reduce the duration of PTS and impair the development of PTE. Acute RTG treatment also prevented post-injury hypersomnia, nerve fiber damage, and cortical TDP-43 accumulation and translocation from the nucleus to the cytoplasm. Mice that developed PTE displayed impaired rapid eye movement (REM) sleep, and there were significant correlations between seizure duration and time spent in the different stages of the sleep-wake cycle. We observed acute RTG treatment to impair injury-induced reduction of age-related increase in gamma frequency power of the EGG, which has been suggested to be necessary for a healthy aged brain. The data show that RTG, administered acutely post-TBI, is a promising, novel therapeutic option to blunt/prevent several long-term effects of rTBIs. Furthermore, our results show a direct relationship between sleep architecture and PTE.
Article
Introduction: Post-mortem analysis provides definitive diagnoses of neurodegenerative diseases; however, only a few can be diagnosed during life. Methods: This study employed statistical tools and machine learning to predict 17 neuropathologic lesions from a cohort of 6518 individuals using 381 clinical features (Table S1). The multisite data allowed validation of the model's robustness by splitting train/test sets by clinical sites. A similar study was performed for predicting Alzheimer's disease (AD) neuropathologic change without specific comorbidities. Results: Prediction results show high performance for certain lesions that match or exceed that of research annotation. Neurodegenerative comorbidities in addition to AD neuropathologic change resulted in compounded, but disproportionate, effects across cognitive domains as the comorbidity number increased. Discussion: Certain clinical features could be strongly associated with multiple neurodegenerative diseases, others were lesion-specific, and some were divergent between lesions. Our approach could benefit clinical research, and genetic and biomarker research by enriching cohorts for desired lesions.
Article
Full-text available
We have previously reported a large Danish pedigree with autosomal dominant frontotemporal dementia (FTD) linked to chromosome 3 (FTD3). Here we identify a mutation in CHMP2B, encoding a component of the endosomal ESCRTIII complex, and show that it results in aberrant mRNA splicing in tissue samples from affected members of this family. We also describe an additional missense mutation in an unrelated individual with FTD. Aberration in the endosomal ESCRTIII complex may result in FTD and neurodegenerative disease.
Article
Full-text available
Objective Familial autosomal dominant frontotemporal dementia with ubiquitin-positive, tau-negative inclusions in the brain linked to 17q21-22 recently has been reported to carry null mutations in the progranulin gene (PGRN). Hereditary dysphasic disinhibition dementia (HDDD) is a frontotemporal dementia with prominent changes in behavior and language deficits. A previous study found significant linkage to chromosome 17 in a HDDD family (HDDD2), but no mutation in the MAPT gene. Longitudinal follow-up has enabled us to identify new cases and to further characterize the dementia in this family. The goals of this study were to develop research criteria to classify the different clinical expressions of dementia observed in this large kindred, to identify the causal mutation in affected individuals and correlate this with phenotypic characteristics in this pedigree, and to assess the neuropathological characteristics using immunohistochemical techniques.Methods In this study we describe a detailed clinical, pathological and mutation analysis of the HDDD2 kindred.ResultsNeuropathologically, HDDD2 represents a familial frontotemporal lobar degeneration with ubiquitin-positive, tau-negative inclusions (FTLD-U). We developed research classification criteria and identified three distinct diagnostic thresholds, which helped localize the disease locus. The chromosomal region with the strongest evidence of linkage lies within the minimum critical region for FTLD-U. Sequencing of each exon of the PGRN gene led to the identification of a novel missense mutation, Ala-9 Asp, within the signal peptide.InterpretationHDDD2 is an FTLD-U caused by a missense mutation in the PGRN gene that cosegregates with the disease and with the disease haplotype in at-risk individuals. This mutation is the first reported pathogenic missense mutation in the signal peptide of the PGRN gene causing FTLD-U. In light of the previous reports of null mutations and its position in the gene, two possible pathological mechanisms are proposed: (1) the protein may accumulate within the endoplasmic reticulum due to inefficient secretion; and (2) mutant RNA may have a lower expression because of degradation via nonsense-mediated decay. Ann Neurol 2006;60:314–322
Article
Early and severe memory impairment is generally held to be an exclusion criterion for the clinical diagnosis of frontotemporal dementia (FTD). However, clinical experience suggests that some patients with otherwise typical FTD can be amnesic from presentation, or even present solely with amnesia. A review of severe amnesia at presentation in patients with pathologically proven FTD is therefore warranted. The present study examined the records of all patients in the joint Cambridge–Sydney neuropathological series of patients with dementia and a pathological diagnosis of FTD to identify those for whom memory complaints were dominant at presentation. Eight of 71 patients met these criteria. For two patients, memory loss was the only complaint; for one patient, memory loss was accompanied by personality change; for two patients, memory loss was accompanied by prominent dysexecutive symptoms; and for three patients, memory loss was accompanied by apathy but no other behavioural changes. In seven patients local specialist teams initially diagnosed Alzheimer's disease; four patients entered anticholinesterase drug trials. All eight later developed behavioural features: in four, the diagnosis was revised to FTD, while in four the diagnosis of FTD was made only on neuropathological examination after death. In conclusion, severe amnesia at presentation in FTD is commoner than previously thought and the clinical consensus criteria for the diagnosis of FTD may need to be revised. The underlying basis of the memory impairments in patients with FTD may be heterogeneous, with different explanations in different subgroups. Abbreviations: FTD = frontotemporal dementia; FTD-MND = frontotemporal dementia with motor neuron inclusions; fvFTD = frontal or behavioural variant frontotemporal dementia; MMSE = Mini-Mental State Examination
Article
To establish survival in patients with pathologically confirmed frontotemporal dementia (FTD) and to determine whether clinical or pathologic subtype affects prognosis. The authors reviewed the presenting clinical features of 61 patients with dementia and pathologically confirmed FTD studied in Sydney (n = 31) and Cambridge (n = 30) over a 10-year period. Data were available on time of symptom onset, diagnosis, institutionalization, and death. Cases were classified pathologically as tau-positive and tau-negative. Of the 61 patients with FTD, 26 presented with frontal variant (fvFTD), 9 with semantic dementia, 8 with progressive nonfluent aphasia (PNFA), 9 with associated motor neuron disease (FTD-MND), and 9 with corticobasal degeneration features. There was no difference between the groups in age at symptom onset (overall mean 58.5 +/- 7.8 years), but at diagnosis the PNFA (68.3 +/- 2.7) group was significantly older than the fvFTD (59.9 +/- 7.4) and FTD-MND (57.7 +/- 7.9) groups. The median survival from symptom onset and from diagnosis was 6 +/- 1.1 years (95% CI) for fvFTD and 3 +/- 0.4 years for FTD-MND. Survival across subgroups was equivalent except for the FTD-MND group, which had significantly shorter survival. Cases with tau-positive pathology had an older age at onset and a significantly better prognosis: median survival 9.0 +/- 0.9 years vs 5.0 +/- 1.1 years. FTD is a malignant disorder with limited life expectancy. FTD-MND has the shortest duration both before and after diagnosis. Tau-positivity is associated with a more slowly progressive form of FTD.
Article
An international group of clinical and basic scientists participated in the Frontotemporal Dementia and Pick's Disease Criteria Conference at the National Institutes of Health in Bethesda, Md, on July 7, 2000, to reassess clinical and neuropathological criteria for the diagnosis of frontotemporal dementia (FTD). Previous criteria for FTD have primarily been designed for research purposes. The goal of this meeting was to propose guidelines that would enable clinicians (particularly neurologists, psychiatrists, and neuropsychologists) to recognize patients with FTD and, if appropriate, to expedite their referral to a diagnostic center. In addition, recommendations for the neuropathological criteria of FTD were reviewed, relative to classical neuropathology and modern molecular biology.