Skip to main content
Brain Pathology logoLink to Brain Pathology
. 2007 Feb 26;17(1):74–82. doi: 10.1111/j.1750-3639.2007.00054.x

Progressive Supranuclear Palsy: Pathology and Genetics

Dennis W Dickson 1, Rosa Rademakers 1, Michael L Hutton 1
PMCID: PMC8095545  PMID: 17493041

Abstract

Progressive supranuclear palsy (PSP) is an atypical Parkinsonian disorder associated with progressive axial rigidity, vertical gaze palsy, dysarthria and dysphagia. Neuropathologically, the subthalamic nucleus and brainstem, especially the midbrain tectum and the superior cerebellar peduncle, show atrophy. The substantia nigra shows loss of pigment corresponding to nigrostriatal dopaminergic degeneration. Microscopic findings include neuronal loss, gliosis and neurofibrillary tangles in basal ganglia, diencephalon and brainstem. Characteristic tau pathology is also found in glia. The major genetic risk factor for sporadic PSP is a common variant in the gene encoding microtubule‐associated protein tau (MAPT) and recent studies have suggested that this may result in the altered expression of specific tau protein isoforms. Imaging studies suggest that there may be sensitive and specific means to differentiate PSP from other parkinsonian disorders, but identification of a diagnostic biomarker is still elusive.

INTRODUCTION

Progressive supranuclear palsy (PSP) is an atypical Parkinsonian disorder associated with progressive axial rigidity, vertical gaze palsy, dysarthria and dysphagia that was first described by Steele, Richardson and Olszewski in 1964 (80). A frontal lobe syndrome and subcortical dementia are sometimes observed (8). Atypical cases may present with apraxia of speech (38), corticobasal syndrome (91) or spastic paraparesis (39). Some patients with pathologically confirmed PSP have a parkinsonian syndrome that is difficult to differentiate from idiopathic Parkinson’s disease, at least initially (95). Given these variants in clinical presentation, it is not surprising that the overall diagnostic accuracy rate for PSP is only 70%–75%(37). In a large autopsy series from the “Society for Progressive Supranuclear Palsy” brain bank at the Mayo Clinic, there are slightly more men than women with PSP (M : F—227:195). The average age at death for cases in the brain bank is 75 ± 8 years and the average disease duration is about 7 years. PSP is a sporadic condition, although about 15% of cases have a family history of neurologic disorders, including parkinsonism and dementia. Kindreds with autosomal‐dominantly inherited PSP have also been described (72).

PATHOLOGY

Macroscopic pathology.  The gross examination of the brain in PSP often reveals distinctive features. Most cases have mild frontal atrophy that may involve the pre‐central gyrus, but the overall brain weight (1170 ± 150 g) may be within normal limits. The midbrain, especially the tectum, and to a lesser extent the pons, show atrophy (Figure 1). The third ventricle and aqueduct of Sylvius may be dilated. The substantia nigra shows loss of pigment, while the locus ceruleus is often better preserved. The subthalamic nucleus is smaller than expected and may have a gray discoloration. The superior cerebellar peduncle and the hilus of the cerebellar dentate nucleus are usually atrophic (90) and have a gray discoloration due to myelinated fiber loss (Figure 2).

Figure 1.

Figure 1

Gross appearance of the brain in progressive supranuclear palsy shows mild frontal atrophy (A) with widening of sulcal spaces (arrows); on medial views (B) the most consistent finding is atrophy of the midbrain with flattening of the tectal plate (arrow).

Figure 2.

Figure 2

Gross appearance of the sectioned brain reveals atrophy of the subthalamic nucleus (arrows) in the diencephalon at the level of the mammillothalamic tract (A), loss of pigment in the substantia nigra (arrow) in the midbrain (B), and atrophy of the superior cerebellar peduncle (arrow) in the rostral pons (C).

Microscopic pathology.  Microscopic findings include neuronal loss, gliosis and neurofibrillary tangles (NFTs) affecting basal ganglia, diencephalon and brainstem. The nuclei most affected are the globus pallidus, subthalamic nucleus and substantia nigra. See Table 1 for distribution of pathology. The cerebral cortex is relatively spared, but lesions are common in the peri‐Rolandic region (30). The limbic lobe is preserved in PSP unless there is concurrent argyrophilic grain disease, which occurs in about 20% of cases (85).

Table 1.

Regional distribution of tau pathology by cell type in PSP.

Region Neurons Astrocytes Oligodendroglia
Cortex Temporal 0.8 0.8 0.2
Motor 1.8 2.6 2.1
Basal ganglia Caudate nucleus 2.0 2.7 1.6
Globus pallidus 2.1 0.8 2.3
Diencephalon Hypothalamus 2.6 0.2 0.4
Basal nucleus 2.8 0.4 0.7
Ventral thalamus 1.9 1.6 2.4
Subthalamic nucleus 2.6 1.3 2.4
Thalamic fasciculus 0.04 2.6
Midbrain Tectum 1.5 2.0 2.4
Oculomotor nucleus 2.2 0.2 0.6
Red nucleus 1.9 1.6 2.5
Substantia nigra 2.4 0.6 1.4
Pons Locus ceruleus 2.8 0.02 0.2
Tegmentum 2.0 0.2 1.9
Base 2.7 0.1 1.0
Medulla Tegmentum 2.6 0.2 1.5
Inferior olive 1.7 0.1 1.0
Cerebellum Dentate nucleus 2.0 0.1 0.8
White matter 0.03 2.2

Average score of 97 cases of PSP; scored on 0–3 point scale for each region and lesion type; darker shading corresponds to more severely affected region. For neurons both pretangles and NFT are scored together; for astrocytes, tufted astrocytes are scored; and for oligodendroglia, coiled bodies are scored.

The striatum and thalamus often have some degree of neuronal loss and gliosis, especially in the ventral anterior and lateral thalamic nuclei. The basal nucleus of Meynert usually has mild cell loss (82). Affected brainstem regions include the superior colliculus, periaqueductal gray matter, oculomotor nuclei, locus ceruleus, pontine nuclei, pontine tegmentum, vestibular nuclei, medullary tegmentum and inferior olives. Cholinergic nuclei in the brainstem appear to be particularly susceptible (94). The cerebellar dentate nucleus is frequently affected and may show grumose degeneration, a type of neuronal degeneration associated with clusters of degenerating presynaptic terminals around dentate neurons (35). The dentatorubrothalamic pathway consistently shows fiber loss. The cerebellar cortex is preserved, but there may be mild Purkinje cell loss and scattered axonal torpedoes. The spinal cord is often affected, with neuronal inclusions in anterior and posterior horns and the intermediolateral cell column 43, 44).

Silver stains (eg, Gallyas stain) or immunostaining for tau reveal NFTs in residual neurons in the basal ganglia, diencephalon, brainstem and spinal cord (Figure 3). In addition to NFTs, special stains demonstrate argyrophilic, tau‐positive inclusions in both astrocytes and oligodendrocytes. Tufted astrocytes are increasingly recognized as a characteristic feature of PSP (56), where they are commonly found in motor cortex and striatum (51). They differ from astrocytic lesions in other neurodegenerative disorders 45, 46). Oligodendroglial changes appear as argyrophilic and tau‐positive perinuclear fibers (3), so‐called “coiled bodies”(9), and are often accompanied by thread‐like processes in white matter, especially in diencephalon and cerebellar white matter.

Figure 3.

Figure 3

The major histologic lesions in progressive supranuclear palsy (PSP) are visible with tau immunohistochemistry (A,C,E) or silver stain (Gallyas) (B,D,F). Neurofibrillary tangles (A,B) often have a globose appearance. Tufted astrocytes (C,D) are relatively specific for PSP. Oligodendroglial coiled bodies (E,F) are characteristic, but can also be found in other disorders.

Neuroinflammation in PSP.  Neuroinflammation occurs in both PSP and the related tauopathy, corticobasal degeneration (CBD), but there are differences in the distribution of microgliosis that to a large extent parallel the tau pathology (35). Activated microglia are abundant in nuclei that show NFTs and neuronal loss, such as the substantia nigra, subthalamic nucleus and globus pallidus, but they are also prominent in certain fiber tracts (eg, dentatorubrothalamic tract), where there is no significant tau pathology. This suggests that degeneration in PSP is linked to tau pathology in a complicated, and sometimes indirect, way.

Electron microscopy.  NFTs in PSP are composed of straight filaments with a diameter of 15 nm that are composed of tau protein 10, 21, 83) (Figure 4). The abnormal filaments in glial cells are also straight and made of tau 3, 56).

Figure 4.

Figure 4

Electron microscopy of neurofibrillary tangle in progressive supranuclear palsy (A) shows a globose tangled mass of filaments that displaces cytosolic elements. The filaments at higher magnification (B) are mostly straight (bars = 1 µm).

Biochemistry.  Biochemical studies show differences between filamentous tau from Alzheimer’s disease (AD) and that from PSP and CBD (10). In AD, detergent‐insoluble tau migrates as three major bands (68, 64 and 60 kDa) on western blots, while in PSP and CBD it migrates as two major bands (68 and 64 kDa) (Figure 5). The 68 and 64 kDa bands of PSP and CBD comprise hyperphosphorylated tau isoforms with four microtubule‐binding repeats (see below). Their contribution to PSP pathology has been confirmed by immunohistochemistry using specific antibodies (50).

Figure 5.

Figure 5

Western blot of detergent insoluble tau proteins in Alzheimer’s disease (AD), corticobasal degeneration (CBD) and two cases of progressive supranuclear palsy (PSP). Note that PSP and CBD have two prominent bands (arrowheads) at 64‐kDa and 68‐kDa, while AD has three major bands (arrowheads) at 60‐kDa, 64‐kDa and 68‐kDa. There is a minor higher molecular weight species in AD, CBD and PSP, as well as variable lower molecular weight species due to protein degradation.

Studies of neurotransmitters reveal loss of striatal dopamine in PSP, as expected for a Parkinsonian disorder (32), as well as deficits in cholinergic markers and basal forebrain cholinergic neurons, but usually not as marked as in AD 36, 82, 94). Other transmitter systems, such as serotonin and gamma‐aminobutyric acid, are affected less consistently (48).

Differential diagnosis.  The major pathologic processes that are clinically mistaken for PSP include multiple system atrophy (MSA), Lewy body disease (LBD) and CBD (37). Neither MSA nor LBD present differential diagnostic difficulties, as they are associated with alpha‐synuclein‐immunoreactive neuronal and glial lesions rather than tau‐immunoreactive deposits. CBD, on the other hand, can present diagnostic difficulties. Table 2 summarizes differential diagnostic features of PSP, CBD and argyrophilic grain disease.

Table 2.

Differential diagnosis of 4R tauopathies. NFT = neurofibrillary tangle.

Feature Corticobasal degeneration Progressive supranuclear palsy Argyrophilic grain disease
Gross appearance Circumscribed atrophy, parasagittal superior frontal and parietal gyri Premotor and frontal atrophy No atrophy or mild medial temporal atrophy
Cortex Superficial spongiosis and gliosis, with rarefaction of white matter Minimal or no histopathology None
Ballooned neurons Abundant in affected cortices None or sparse, limbic lobe Sparse, limbic lobe
Tau immunoreactive neuronal lesions Pleomorphic skein‐like inclusions, pretangles, and sparse NFTs Globose NFTs, pretangles Pretangles and grains
Threads and thread‐like lesions Numerous, gray and white matter cortical and striatal Numerous in diencephalon and brainstem tectum and tegmentum Minimal, medial temporal lobe
Glial pathology Astrocytic plaques, coiled bodies Tufted astrocytes, coiled bodies Coiled bodies, tau‐positive ramified astrocytes
Subthalamic nucleus Variable gliosis and pretangles, but minimal neuronal loss Marked neuronal loss and gliosis with neuronal and glial tau pathology Minimal or no pathology
Substantia nigra Marked neuronal loss and gliosis Marked neuronal loss and gliosis None
Brainstem Minimal or no neuronal loss, but pretangles and threads are present in certain nuclei (eg, locus ceruleus) Variable neuronal loss (eg, red nucleus) and many pretangles and threads None
Cerebellum Minimal tau pathology, mostly in dentate nucleus; dentate grumose degeneration is uncommon Tau pathology in dentate nucleus and white matter; dentate grumose degeneration is common None

Clinically, both PSP and CBD give rise to extrapyramidal signs similar to Parkinson’s disease, but neither condition is responsive to levodopa therapy. Motor abnormalities in PSP are usually symmetrical, while asymmetry is the hallmark of CBD. Focal cortical signs, such as apraxia and aphasia, are common in CBD, but rare in PSP. Frontal dementia is more common in CBD than in PSP (7).

Pathologically, both PSP and CBD are associated with neuronal and glial filamentous lesions composed of hyperphosphorylated four‐repeat tau protein. The morphologies of the lesions differ (20). In CBD only a minority of the neuronal lesions consists of well‐formed NFTs. Most lesions are pleomorphic filamentous inclusions. Nerve cells can also display diffuse granular cytoplasmic tau immunoreactivity that is non‐argyrophilic, consistent with so‐called “pretangles”. In contrast, globose and argyrophilic NFTs are more common in PSP.

The astrocytic lesions are also different. In CBD the characteristic lesion is the so‐called astrocytic plaque (25), which is characterized by clusters of short and irregular tau‐immunoreactive cell processes around the unstained cell body. In contrast, the tufted astrocyte of PSP has tau‐immunoreactive cell processes that extend to the stained cell body. The most characteristic lesions of CBD are thread‐like tau‐immunoreactive processes in both gray and white matter of cerebrum, diencephalon and rostral brainstem (23). In CBD there are usually only a few oligodendroglial coiled bodies. They are more numerous in PSP, especially in areas vulnerable to thread‐like pathology. Thread‐like processes in PSP have a different distribution from those in CBD, being sparse in cerebrum, but dense in diencephalon and brainstem, as well as in cerebellar white matter (26). Both diseases affect the cerebellar dentate nucleus, with grumose degeneration being much more common in PSP (34) than in CBD (71).

Another characteristic feature of CBD are the swollen, achromatic, or ballooned neurons (71). They correspond to swollen pyramidal cells, in affected cortical areas (eg, superior frontal gyrus), that are immunoreactive for neurofilaments and alpha‐B‐crystallin 22, 42). They are not typical of PSP; however, when present, it is usually in the setting of concurrent argyrophilic grain disease (84).

GENETICS

The tau gene (MAPT).  PSP and CBD are considered non‐familial or sporadic tauopathies, although rare multiplex kindreds have been described. A genetic basis for familial clustering has been demonstrated in a Spanish family with autosomal‐dominant PSP that is linked to chromosome 1q31 72, 73, 97) and in families with causal mutations in the microtubule‐associated protein tau gene (MAPT) (68). In addition, numerous studies have identified common genetic variability in MAPT as a major risk factor for PSP.

MAPT is located on the long arm of chromosome 17 (17q21.1) and consists of a non‐coding exon 0 followed by 14 coding or alternatively spliced exons 1, 68). In humans, MAPT exons 2, 3, 4A, 6, 8 and 10 undergo tissue‐specific and developmentally regulated alternative mRNA splicing. In brain, MAPT exons 4A, 6 and 8 are not usually transcribed; however, alternative splicing of MAPT exons 2, 3 and 10 gives rise to six major tau isoforms composed of 352–441 amino acids 1, 29). Binding to microtubules occurs through the C‐terminal half, where tau contains four imperfect repeats that are encoded by exons 9–12, with alternative mRNA splicing of exon 10 giving rise to classes of tau isoforms with either three (3R) or four (4R) repeats (29). In vitro studies have shown that 4R tau binds more strongly to microtubules than 3R tau (61). The isoform composition in brain is developmentally regulated, with only 3R tau being expressed in fetal brain and similar levels of 3R and 4R tau being found in adult human brain (29). In neurodegenerative disorders, the isoform composition of pathologic tau varies in predictable ways. Neurofibrillary pathology in Alzheimer’s disease is composed of hyperphosphorylated of 3R and 4R tau, while Pick bodies of Pick’s disease contain predominantly hyperphosphorylated 3R tau (10). The family of 4R tauopathies includes PSP and CBD, as well as the age‐related medial temporal lobe tauopathy, argyrophilic grain disease (86).

Frontotemporal dementia and Parkinsonism linked to chromosome 17 (FTDP‐17) may be associated with mutations in MAPT 33, 66, 77). To date, 39 different pathogenic mutations have been reported in 115 families worldwide (68). In some families, affected individuals have clinical and pathologic features that overlap with those of PSP and CBD 11, 13, 19, 54, 62, 67, 74, 76, 78, 79). At least three MAPT mutations have been shown to cause a pathologically confirmed PSP phenotype: R5L in exon 1, ΔN296 in exon 10 and G303V in exon 10 67, 74, 76, 79).

Before it was known that MAPT mutations lead to neurodegeneration in FTDP‐17, Conrad et al had reported that common genetic variability in MAPT could lead to increased risk for PSP 15, 16). The initial association study was performed with a dinucleotide repeat polymorphism in intron 9 of MAPT, presenting with four polymorphic alleles, two of which, A0 and A3, are common. Significant overrepresentation of the A0 allele was observed in PSP patients compared with controls, a finding replicated in at least four independent PSP populations 6, 31, 53, 59). More recently, it was recognized that MAPT is located in a complex genomic region that is surrounded by three highly homologous low‐copy repeats 17, 81). About 3 million years ago, these low‐copy repeats appear to have induced a 900 kb genetically balanced inversion that resulted in a high degree of linkage disequilibrium in the MAPT genomic region, producing the two extended MAPT haplotypes H1 and H2 (4). The A0 allele of the dinucleotide repeat in intron 9 is inherited as part of the H1 haplotype, resulting in a consistent association of PSP with the extended H1 haplotype in all analyzed Caucasian populations (4). In other populations, for example the Japanese, virtually all individuals are homozygous for the H1 haplotype 16, 24).

The sequence of tau protein is the same on the H1 and H2 backgrounds, suggesting that the association of MAPT with PSP relates to differences in tau expression levels, in alternative mRNA splicing, or in a combination of both. As H1 is common in the general population, it is expected that subtle genetic differences on this haplotype, rather than the haplotype per se, are responsible for the increased PSP risk.

Using high‐density association studies with a panel of H1‐specific tagging single nucleotide polymorphisms (SNPs), it was shown that the PSP risk could be explained in part by one of two major ancestral H1 haplotypes, named H1b by Rademakers et al (69) and H1c by Pittman et al (65). Analyses in young PSP patients localized the H1b/c risk to a 22 kb regulatory region in MAPT intron 0, defined by one SNP (rs242557) located in a highly conserved sequence element (69). In vitro assays of gene expression have indicated that this conserved region is indeed transcriptionally active, with both alleles of rs242557 differentially influencing expression. The functional relevance of the H1b/c haplotype defined by the “A” allele of rs242557 has been confirmed by extensive reporter gene analyses, which have shown that H1b/c increases levels of all MAPT transcripts and, in particular, those encoding 4R tau (55). A recent study focusing on allele‐specific gene expression in H1/H2 heterozygous neuronal cell cultures and post‐mortem human brain tissue has reported that the level of MAPT expression from H1 and H2 was not significantly different, but that about 1.5‐fold more 4R tau‐containing transcripts were expressed from H1 (12). The implication of these genetic studies is that the risk for PSP may be associated with lifelong higher levels of 4R tau expression in the nervous system.

Other genes.  One of the most common genetic causes of Parkinson’s disease is mutations in a gene on chromosome 12, LRRK2 60, 98). While most individuals with LRRK2 mutations have typical Lewy body Parkinson’s disease, there are individuals in some kindreds that have primarily a tau pathology 70, 96, 98). The lesions in these cases bear resemblance to those found in PSP, but they differ in the density and distribution. Consistent with these observations is the fact that mutations have not been found in LRRK2 in a large series of pathologically confirmed PSP (75).

Recently, the genetic basis for a major form of frontotemporal lobar degeneration associated with ubiquitin‐immunoreactive neuronal inclusions has been discovered 5, 18). Progranulin (PRGN), the mutant gene, is located close to MAPT on chromosome 17 5, 18, 27). Interestingly, while the most common clinical presentation of PGRN mutation carriers is frontal lobe dementia with progressive language failure, more than half of the pathologically confirmed cases had Parkinsonism and several cases came to autopsy with an initial clinical diagnosis of PSP (41). The pathology is distinctly different from PSP and is associated with cortical, limbic and striatal degeneration with neuronal cytoplasmic and intranuclear inclusions. NFTs and astrocytic lesions are not detected. To date, no case of PSP has been shown to be associated with a PGRN mutation.

Mixed pathology.  Given the advanced age of many patients with PSP, it is not surprising that they should have some degree of Alzheimer‐type pathology. In most cases this is minor, but rare individuals have sufficient senile plaques and NFTs to warrant a diagnosis of concurrent AD. The risk factors for AD in the setting of PSP are advanced age, female sex and carrier status of apolipoprotein E4 (89).

Lewy bodies are also detected in some cases of PSP, but in a distribution that is similar to that in incidental LBD or Parkinson’s disease. The frequency of Lewy bodies in PSP is about 10%, which is similar to the frequency in normal elderly controls 88, 92).

The environmental risk factors for PSP are unknown (49), but some studies have suggested that hypertension may be more frequent in PSP (28). However, other studies have failed to confirm this association (14). There is also no significant increase in the frequency of arteriosclerotic small vessel disease or lacunar infarcts in PSP compared with controls, which might be expected if hypertension was a risk factor for PSP.

Argyrophilic grain disease, like PSP and CBD, is a 4R tauopathy (86). It is a pathologic process in which tau accumulates in dendrites of pyramidal neurons as grain‐like structures. The lesions are concentrated in the medial temporal lobe. Ballooned neurons in argyrophilic grain disease are found in limbic areas, including the amygdala, entorhinal cortex, cingulate gyrus and claustrum (87). These are areas where ballooned neurons are also detected in PSP and almost all cases of PSP with ballooned neurons have a medial temporal lobe 4R tauopathy similar to argyrophilic grain disease.

Atypical cases.  While dementia occurs in some patients with PSP, it is usually relatively mild and characterized by a subcortical frontal‐executive syndrome, featuring cognitive slowing, poor memory retrieval and poor planning and organization. On the other hand, some patients have prominent cortical signs and features suggestive of a frontal lobe syndrome. The frontal lobe degeneration syndromes include frontal lobe dementia with prominent personality changes (apathy or disinhibition) (8), progressive language failure (progressive aphasia or semantic dementia) (38) and progressive asymmetrical rigidity and apraxia (corticobasal syndrome) (91). Some cases with autopsy‐proven PSP‐type pathology have each of these syndromes. As a general rule, when PSP presents with one of these syndromes, the cortical pathology extends beyond the peri‐Rolandic cortices to frontal, parietal or peri‐Sylvian areas. Morphologically, however, the lesions are similar to those in typical PSP.

Motor neuron degeneration is unusual in PSP, but some patients have been described to have long tract signs (eg, hyperreflexia, Babinski signs and clonus). Less common are cases with features suggestive of primary lateral sclerosis (39). These patients present diagnostic difficulties and are often considered to have frontal lobe degeneration with motor neuron disease or CBD. At autopsy, they have pathology similar to typical PSP, but more marked neuronal loss and gliosis in motor cortex, including loss of Betz cells (38). There is also degeneration of the corticospinal tract that can be traced from motor cortex to lower brainstem regions. Most of these cases have less pathology in the subthalamic nucleus and substantia nigra than typical cases of PSP. Histologically, the neuronal and glial lesions are similar to those in PSP, but some cases have peculiar globular inclusions in oligodendroglial cells.

At autopsy, the pathology of PSP is often found to underlie the clinical syndrome of pure akinesia 47, 52), which is characterized by bradykinesia with gait freezing. These patients do usually not suffer early falls or eye movement abnormalities. Thus, several cardinal clinical features of PSP are not met. Pathological changes are often milder and more concentrated in brainstem and diencephalon, with fewer changes in cortex and striatum.

Recently, Williams et al have drawn attention to the fact that some patients with PSP can present with an asymmetrical extrapyramidal syndrome with tremor and a positive response to levodopa therapy. They referred to this syndrome as PSP‐P (95). Most of these patients developed eventually the clinical features of PSP, but early in the disease their symptoms resembled those of Parkinson’s disease. The term Richardson’s syndrome was proposed to refer to the more typical PSP syndrome (95). The pathology of PSP‐P is virtually indistinguishable from that of PSP, except that cortex and striatum are less severely affected.

Biomarkers in PSP.  A definitive cerebrospinal fluid biomarker for PSP remains elusive. Several studies have measured CSF tau protein and found it to be variably increased in PSP, but not more than in AD or CBD 2, 57, 93). Thus, it cannot be used to differentiate PSP from these disorders.

Structural imaging holds promise for differentiating PSP from other parkinsonian disorders. In particular, indices of brainstem pathology, such as atrophy of the superior cerebellar peduncle (90), are present consistently and differ from the changes observed in idiopathic Parkinson’s disease and MSA 58, 63). Longitudinal measures also show a more marked cortical and brainstem atrophy in PSP, particularly of the midbrain, than in normal controls 40, 64).

ACKNOWLEDGMENTS

The contributions from the Society for Progressive Supranuclear Palsy brain bank were crucial to this article. The authors also acknowledge Zeshan Ahmed for Gallyas silver stains, Wen‐Lang Lin for electron microscopy and Shu‐Hui Yen for western blot studies.

This study was supported by: P50‐AG25711, P50‐AG16574, P50‐NS40256, P01‐AG17216, P01‐AG03949, P01‐AG14449, R01‐ES013941, R01‐AG15866 and the Society for Progressive Supranuclear Palsy.

REFERENCES

  • 1. Andreadis A, Brown WM, Kosik KS (1992) Structure and novel exons of the human tau gene. Biochemistry 31:10626–10633. [DOI] [PubMed] [Google Scholar]
  • 2. Arai H, Morikawa Y, Higuchi M, Matsui T, Clark CM, Miura M, Machida N, Lee VM, Trojanowski JQ, Sasaki H (1997) Cerebrospinal fluid tau levels in neurodegenerative diseases with distinct tau‐related pathology. Biochem Biophys Res Commun 236:262–264. [DOI] [PubMed] [Google Scholar]
  • 3. Arima K, Nakamura M, Sunohara N, Ogawa M, Anno M, Izumiyama Y, Hirai S, Ikeda K (1997) Ultrastructural characterization of the tau‐immunoreactive tubules in the oligodendroglial perikarya and their inner loop processes in progressive supranuclear palsy. Acta Neuropathol (Berl) 93:558–566. [DOI] [PubMed] [Google Scholar]
  • 4. Baker M, Litvan I, Houlden H, Adamson J, Dickson D, Perez‐Tur J, Hardy J, Lynch T, Bigio E, Hutton M (1999) Association of an extended haplotype in the tau gene with progressive supranuclear palsy. Hum Mol Genet 8:711–715. [DOI] [PubMed] [Google Scholar]
  • 5. Baker M, Mackenzie IR, Pickering‐Brown SM, Gass J, Rademakers R, Lindholm C, Snowden J, Adamson J, Sadovnick AD, Rollinson S, Cannon A, Dwosh E, Neary D, Melquist S, Richardson A, Dickson D, Berger Z, Eriksen J, Robinson T, Zehr C, Dickey CA, Crook R, McGowan E, Mann D, Boeve B, Feldman H, Hutton M (2006) Mutations in progranulin cause tau‐negative frontotemporal dementia linked to chromosome 17. Nature 442: 916–919. [DOI] [PubMed] [Google Scholar]
  • 6. Bennett P, Bonifati V, Bonuccelli U, Colosimo C, De Mari M, Fabbrini G, Marconi R, Meco G, Nicholl DJ, Stocchi F, Vanacore N, Vieregge P, Williams AC (1998) Direct genetic evidence for involvement of tau in progressive supranuclear palsy. European Study Group on Atypical Parkinsonism Consortium. Neurology 51:982–985. [DOI] [PubMed] [Google Scholar]
  • 7. Bergeron C, Davis A, Lang AE (1998) Corticobasal ganglionic degeneration and progressive supranuclear palsy presenting with cognitive decline. Brain Pathol 8:355–365. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 8. Bigio EH, Brown DF, White CL 3rd (1999) Progressive supranuclear palsy with dementia: cortical pathology. J Neuropathol Exp Neurol 58:359–364. [DOI] [PubMed] [Google Scholar]
  • 9. Braak H, Braak E (1989) Cortical and subcortical argyrophilic grains characterize a disease associated with adult onset dementia. Neuropathol Appl Neurobiol 15:13–26. [DOI] [PubMed] [Google Scholar]
  • 10. Buée L, Delacourte A (1999) Comparative biochemistry of tau in progressive supranuclear palsy, corticobasal degeneration, FTDP‐17 and Pick’s disease. Brain Pathol 9:681–693. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 11. Bugiani O, Murrell JR, Giaccone G, Hasegawa M, Ghigo G, Tabaton M, Morbin M, Primavera A, Carella F, Solaro C, Grisoli M, Savoiardo M, Spillantini MG, Tagliavini F, Goedert M, Ghetti B (1999) Frontotemporal dementia and corticobasal degeneration in a family with a P301S mutation in tau. J Neuropathol Exp Neurol 58:667–677. [DOI] [PubMed] [Google Scholar]
  • 12. Caffrey TM, Joachim C, Paracchini S, Esiri MM, Wade‐Martins R (2006) Haplotype‐specific expression of exon 10 at the human MAPT locus. Hum Mol Genet 15:3529–3537. [DOI] [PubMed] [Google Scholar]
  • 13. Casseron W, Azulay JP, Guedj E, Gastaut JL, Pouget J (2005) Familial autosomal dominant corticobasal degeneration with the P301S mutation in the tau gene: an example of phenotype variability. J Neurol 252:1546–1548. [DOI] [PubMed] [Google Scholar]
  • 14. Colosimo C, Osaki Y, Vanacore N, Lees AJ (2003) Lack of association between progressive supranuclear palsy and arterial hypertension: a clinicopathological study. Mov Disord 18:694–697. [DOI] [PubMed] [Google Scholar]
  • 15. Conrad C, Andreadis A, Trojanowski JQ, Dickson DW, Kang D, Chen X, Wiederholt W, Hansen L, Masliah E, Thal LJ, Katzman R, Xia Y, Saitoh T (1997) Genetic evidence for the involvement of tau in progressive supranuclear palsy. Ann Neurol 41: 277–281. [DOI] [PubMed] [Google Scholar]
  • 16. Conrad C, Amano N, Andreadis A, Xia Y, Namekataf K, Oyama F, Ikeda K, Wakabayashi K, Takahashi H, Thal LJ, Katzman R, Shackelford DA, Matsushita M, Masliah E, Sawa A (1998) Differences in a dinucleotide repeat polymorphism in the tau gene between Caucasian and Japanese populations: implication for progressive supranuclear palsy. Neurosci Lett 250:135–137. [DOI] [PubMed] [Google Scholar]
  • 17. Cruts M, Rademakers R, Gijselinck I, Van Der Zee J, Dermaut B, De Pooter T, De Rijk P, Del‐Favero J, Van Broeckhoven C (2005) Genomic architecture of human 17q21 linked to frontotemporal dementia uncovers a highly homologous family of low‐copy repeats in the tau region. Hum Mol Genet 14:1753–1762. [DOI] [PubMed] [Google Scholar]
  • 18. Cruts M, Gijselinck I, Van Der Zee J, Engelborghs S, Wils H, Pirici D, Rademakers R, Vandenberghe R, Dermaut B, Martin JJ, Van Duijn C, Peeters K, Sciot R, Santens P, De Pooter T, Mattheijssens M, Van den Broeck M, Cuijt I, Vennekens K, De Deyn PP, Kumar‐Singh S, Van Broeckhoven C (2006) Null mutations in progranulin cause ubiquitin‐positive frontotemporal dementia linked to chromosome 17q21. Nature 442:920–924. [DOI] [PubMed] [Google Scholar]
  • 19. Delisle MB, Murrell JR, Richardson R, Trofatter JA, Rascol O, Soulages X, Mohr M, Calvas P, Ghetti B (1999) A mutation at codon 279 (N279K) in exon 10 of the Tau gene causes a tauopathy with dementia and supranuclear palsy. Acta Neuropathol 98:62–77. [DOI] [PubMed] [Google Scholar]
  • 20. Dickson DW (1999) Neuropathologic differentiation of progressive supranuclear palsy and corticobasal degeneration. J Neurol 246(Suppl. 2): II6–II15. [DOI] [PubMed] [Google Scholar]
  • 21. Dickson DW, Kress Y, Crowe A, Yen SH (1985) Monoclonal antibodies to Alzheimer neurofibrillary tangles. 2. Demonstration of a common antigenic determinant between ANT and neurofibrillary degeneration in progressive supranuclear palsy. Am J Pathol 120:292–303. [PMC free article] [PubMed] [Google Scholar]
  • 22. Dickson DW, Yen SH, Suzuki KI, Davies P, Garcia JH, Hirano A (1986) Ballooned neurons in select neurodegenerative diseases contain phosphorylated neurofilament epitopes. Acta Neuropathol (Berl) 71:216–223. [DOI] [PubMed] [Google Scholar]
  • 23. Dickson DW, Bergeron C, Chin SS, Duyckaerts C, Horoupian D, Ikeda K, Jellinger K, Lantos PL, Lippa CF, Mirra SS, Tabaton M, Vonsattel JP, Wakabayashi K, Litvan I (2002) Office of Rare Diseases neuropathologic criteria for corticobasal degeneration. J Neuropathol Exp Neurol 61:935–946. [DOI] [PubMed] [Google Scholar]
  • 24. Evans W, Fung HC, Steele J, Eerola J, Tienari P, Pittman A, Silva R, Myers A, Vrieze FW, Singleton A, Hardy J (2004) The tau H2 haplotype is almost exclusively Caucasian in origin. Neurosci Lett 369:183–185. [DOI] [PubMed] [Google Scholar]
  • 25. Feany MB, Dickson DW (1995) Widespread cytoskeletal pathology characterizes corticobasal degeneration. Am J Pathol 146:1388–1396. [PMC free article] [PubMed] [Google Scholar]
  • 26. Feany MB, Mattiace LA, Dickson DW (1996) Neuropathologic overlap of progressive supranuclear palsy, Pick’s disease and corticobasal degeneration. J Neuropathol Exp Neurol 55:53–67. [DOI] [PubMed] [Google Scholar]
  • 27. Gass J, Cannon A, Mackenzie IR, Boeve B, Baker M, Adamson J, Crook R, Melquist S, Kuntz K, Petersen R, Josephs K, Pickering‐Brown SM, Graff‐Radford N, Uitti R, Dickson D, Wszolek Z, Gonzalez J, Beach TG, Bigio E, Johnson N, Weintraub S, Mesulam M, White CL 3rd, Woodruff B, Caselli R, Hsiung GY, Feldman H, Knopman D, Hutton M, Rademakers R (2006) Mutations in progranulin are a major cause of ubiquitin‐positive frontotemporal lobar degeneration. Hum Mol Genet 15:2988–3001. [DOI] [PubMed] [Google Scholar]
  • 28. Ghika J, Bogousslavsky J (1997) Presymptomatic hypertension is a major feature in the diagnosis of progressive supranuclear palsy. Arch Neurol 54:1104–1108. [DOI] [PubMed] [Google Scholar]
  • 29. Goedert M, Spillantini MG, Jakes R, Rutherford D, Crowther RA (1989) Multiple isoforms of human microtubule‐associated protein tau: sequences and localization in neurofibrillary tangles of Alzheimer’s disease. Neuron 3:519–526. [DOI] [PubMed] [Google Scholar]
  • 30. Hauw JJ, Verny M, Delaère P, Cervera P, He Y, Duyckaerts C (1990) Constant neurofibrillary changes in the neocortex in progressive supranuclear palsy. Basic differences with Alzheimer’s disease and aging. Neurosci Lett 119:182–186. [DOI] [PubMed] [Google Scholar]
  • 31. Higgins JJ, Litvan I, Pho LT, Li W, Nee LE (1998) Progressive supranuclear gaze palsy is in linkage disequilibrium with the tau and not the alpha‐synuclein gene. Neurology 50:270–273. [DOI] [PubMed] [Google Scholar]
  • 32. Hornykiewicz O, Shannak K (1994) Brain monoamines in progressive supranuclear palsy–comparison with idiopathic Parkinson’s disease. J Neural Transm Suppl 42:219–227. [DOI] [PubMed] [Google Scholar]
  • 33. Hutton M, Lendon CL, Rizzu P, Baker M, Froelich S, Houlden H, Pickering‐Brown S, Chakraverty S, Isaacs A, Grover A, Hackett J, Adamson J, Lincoln S, Dickson D, Davies P, Petersen RC, Stevens M, De Graaff E, Wauters E, Van Baren J, Hillebrand M, Joosse M, Kwon JM, Nowotny P, Che LK, Norton J, Morris JC, Reed LA, Trojanowski J, Basun H, Lannfelt L, Neystat M, Fahn S, Dark F, Tannenberg T, Dodd PR, Hayward N, Kwok JB, Schofield PR, Andreadis A, Snowden J, Craufurd D, Neary D, Owen F, Oostra BA, Hardy J, Goate A, Van Swieten J, Mann D, Lynch T, Heutink P (1998) Association of missense and 5′‐splice‐site mutations in tau with the inherited dementia FTDP‐17. Nature 393:702–705. [DOI] [PubMed] [Google Scholar]
  • 34. Ishizawa K, Dickson DW (2001) Microglial activation parallels system degeneration in progressive supranuclear palsy and corticobasal degeneration. J Neuropathol Exp Neurol 60:647–657. [DOI] [PubMed] [Google Scholar]
  • 35. Ishizawa K, Lin WL, Tiseo P, Honer WG, Davies P, Dickson DW (2000) A qualitative and quantitative study of grumose degeneration in progressive supranuclear palsy. J Neuropathol Exp Neurol 59:513–524. [DOI] [PubMed] [Google Scholar]
  • 36. Javoy‐Agid F (1994) Cholinergic and peptidergic systems in PSP. J Neural Transm Suppl 42:205–218. [DOI] [PubMed] [Google Scholar]
  • 37. Josephs KA, Dickson DW (2003) Diagnostic accuracy of progressive supranuclear palsy in the Society for Progressive Supranuclear Palsy brain bank. Mov Disord 18:1018–1026. [DOI] [PubMed] [Google Scholar]
  • 38. Josephs KA, Boeve BF, Duffy JR, Smith GE, Knopman DS, Parisi JE, Petersen RC, Dickson DW (2005) Atypical progressive supranuclear palsy underlying progressive apraxia of speech and nonfluent aphasia. Neurocase 11:283–296. [DOI] [PubMed] [Google Scholar]
  • 39. Josephs KA, Katsuse O, Beccano‐Kelly DA, Lin WL, Uitti RJ, Fujino Y, Boeve BF, Hutton ML, Baker MC, Dickson DW (2006) Atypical progressive supranuclear palsy with corticospinal tract degeneration. J Neuropathol Exp Neurol 65:396–405. [DOI] [PubMed] [Google Scholar]
  • 40. Josephs KA, Whitwell JL, Boeve BF, Shiung MM, Gunter JL, Parisi JE, Dickson DW, Jack CR (2006) Rates of cerebral atrophy in autopsy‐confirmed progressive supranuclear palsy. Ann Neurol 59: 200–203. [DOI] [PubMed] [Google Scholar]
  • 41. Josephs KA, Ahmed Z, Katsuse O, Parisi JF, Boeve BF, Knopman DS, Petersen RC, Davies P, Duara R, Graff‐Radford NR, Uitti RJ, Rademakers R, Adamson J, Baker M, Hutton ML, Dickson DW (2007) Neuropathologic features of frontotemporal lobar degeneration with ubiquitin‐positive inclusions with progranulin gene (PGRN) mutations. J Neuropathol Exp Neurol 66:142–151. [DOI] [PubMed] [Google Scholar]
  • 42. Kato S, Hirano A, Umahara T, Kato M, Herz F, Ohama E (1992) Comparative immunohistochemical study on the expression of alpha B crystallin, ubiquitin and stress‐response protein 27 in ballooned neurons in various disorders. Neuropathol Appl Neurobiol 18:335–340. [DOI] [PubMed] [Google Scholar]
  • 43. Kato T, Hirano A, Weinberg MN, Jacobs AK (1986) Spinal cord lesions in progressive supranuclear palsy: some new observations. Acta Neuropathol (Berl) 71:11–14. [DOI] [PubMed] [Google Scholar]
  • 44. Kikuchi H, Doh‐ura K, Kira J, Iwaki T (1999) Preferential neurodegeneration in the cervical spinal cord of progressive supranuclear palsy. Acta Neuropathol (Berl) 97:577–584. [DOI] [PubMed] [Google Scholar]
  • 45. Komori T (1999) Tau‐positive glial inclusions in progressive supranuclear palsy, corticobasal degeneration and Pick’s disease. Brain Pathol 9:663–679. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46. Komori T, Arai N, Oda M, Nakayama H, Mori H, Yagishita S, Takahashi T, Amano N, Murayama S, Murakami S, Shibata N, Kobayashi M, Sasaki S, Iwata M (1998) Astrocytic plaques and tufts of abnormal fibers do not coexist in corticobasal degeneration and progressive supranuclear palsy. Acta Neuropathol (Berl) 96:401–408. [DOI] [PubMed] [Google Scholar]
  • 47. Konishi Y, Shirabe T, Katayama S, Funakawa I, Terao A (2005) Autopsy case of pure akinesia showing pallidonigro‐luysian atrophy. Neuropathology 25:220–227. [DOI] [PubMed] [Google Scholar]
  • 48. Landwehrmeyer B, Palacios JM (1994) Alterations of neurotransmitter receptors and neurotransmitter transporters in progressive supranuclear palsy. J Neural Transm Suppl 42:229–246. [DOI] [PubMed] [Google Scholar]
  • 49. Litvan I (2003) Update on epidemiological aspects of progressive supranuclear palsy. Mov Disord 18(Suppl. 6):S43–S50. [DOI] [PubMed] [Google Scholar]
  • 50. Liu WK, Le TV, Adamson J, Baker M, Cookson N, Hardy J, Hutton M, Yen SH, Dickson DW (2001) Relationship of the extended tau haplotype to tau biochemistry and neuropathology in progressive supranuclear palsy. Ann Neurol 50:494–502. [DOI] [PubMed] [Google Scholar]
  • 51. Matsusaka H, Ikeda K, Akiyama H, Arai T, Inoue M, Yagishita S (1998) Astrocytic pathology in progressive supranuclear palsy: significance for neuropathological diagnosis. Acta Neuropathol (Berl) 96:248–252. [DOI] [PubMed] [Google Scholar]
  • 52. Mizusawa H, Mochizuki A, Ohkoshi N, Yoshizawa K, Kanazawa I, Imai H (1993) Progressive supranuclear palsy presenting with pure akinesia. Adv Neurol 60:618–621. [PubMed] [Google Scholar]
  • 53. Morris HR, Janssen JC, Bandmann O, Daniel SE, Rossor MN, Lees AJ, Wood NW (1999) The tau gene A0 polymorphism in progressive supranuclear palsy and related neurodegenerative diseases. J Neurol Neurosurg Psychiatry 66:665–667. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 54. Morris HR, Osaki Y, Lees AJ, Wood NW, Revesz T, Quinn N (2003) Tau exon 10+16 mutation FTDP‐17 presenting clinically as sporadic young onset PSP. Neurology 61:102–104. [DOI] [PubMed] [Google Scholar]
  • 55. Myers AJ, Pittman AM, Zhao AS, Rohrer K, Kaleem M, Marlowe L, Lees A, Leung D, McKeith IG, Perry RH, Morris CM, Trojanowski JQ, Clark C, Karlawish J, Arnold S, Forman MS, Deerlin VV, De Silva R, Hardy J (2007) The MAPT H1c risk haplotype is associated with incresed expression of tau and especially of 4 repeat containing transcripts. Neurobiol Dis (in press). [DOI] [PubMed] [Google Scholar]
  • 56. Nishimura M, Namba Y, Ikeda K, Oda M (1992) Glial fibrillary tangles with straight tubules in the brains of patients with progressive supranuclear palsy. Neurosci Lett 143:35–38. [DOI] [PubMed] [Google Scholar]
  • 57. Noguchi M, Yoshita M, Matsumoto Y, Ono K, Iwasa K, Yamada M (2005) Decreased beta‐amyloid peptide42 in cerebrospinal fluid of patients with progressive supranuclear palsy and corticobasal degeneration. J Neurol Sci 237:61–65. [DOI] [PubMed] [Google Scholar]
  • 58. Oba H, Yagishita A, Terada H, Barkovich AJ, Kutomi K, Yamauchi T, Furui S, Shimizu T, Uchigata M, Matsumura K, Sonoo M, Sakai M, Takada K, Harasawa A, Takeshita K, Kohtake H, Tanaka H, Suzuki S (2005) New and reliable MRI diagnosis for progressive supranuclear palsy. Neurology 64:2050–2055. [DOI] [PubMed] [Google Scholar]
  • 59. Oliva R, Tolosa E, Ezquerra M, Molinuevo JL, Valldeoriola F, Burguera J, Calopa M, Villa M, Ballesta F (1998) Significant changes in the tau A0 and A3 alleles in progressive supranuclear palsy and improved genotyping by silver detection. Arch Neurol 55:1122–1124. [DOI] [PubMed] [Google Scholar]
  • 60. Paisan‐Ruiz C, Jain S, Evans EW, Gilks WP, Simon J, Van Der Brug M, Lopez de Munain A, Aparicio S, Gil AM, Khan N, Johnson J, Martinez JR, Nicholl D, Carrera IM, Pena AS, De Silva R, Lees A, Marti‐Masso JF, Perez‐Tur J, Wood NW, Singleton AB (2004) Cloning of the gene containing mutations that cause PARK8‐linked Parkinson’s disease. Neuron 44:595–600. [DOI] [PubMed] [Google Scholar]
  • 61. Panda D, Samuel JC, Massie M, Feinstein SC, Wilson L (2003) Differential regulation of microtubule dynamics by three‐ and four‐repeat tau: implications for the onset of neurodegenerative disease. Proc Natl Acad Sci USA 100:9548–9553. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 62. Pastor P, Pastor E, Carnero C, Vela R, Garcia T, Amer G, Tolosa E, Oliva R (2001) Familial atypical progressive supranuclear palsy associated with homozigosity for the delN296 mutation in the tau gene. Ann Neurol 49:263–267. [DOI] [PubMed] [Google Scholar]
  • 63. Paviour DC, Price SL, Stevens JM, Lees AJ, Fox NC (2005) Quantitative MRI measurement of superior cerebellar peduncle in progressive supranuclear palsy. Neurology 64:675–679. [DOI] [PubMed] [Google Scholar]
  • 64. Paviour DC, Price SL, Jahanshahi M, Lees AJ, Fox NC (2006) Longitudinal MRI in progressive supranuclear palsy and multiple system atrophy: rates and regions of atrophy. Brain 129:1040–1049. [DOI] [PubMed] [Google Scholar]
  • 65. Pittman AM, Myers AJ, Abou‐Sleiman P, Fung HC, Kaleem M, Marlowe L, Duckworth J, Leung D, Williams D, Kilford L, Thomas N, Morris CM, Dickson D, Wood NW, Hardy J, Lees AJ, De Silva R (2005) Linkage disequilibrium fine mapping and haplotype association analysis of the tau gene in progressive supranuclear palsy and corticobasal degeneration. J Med Genet 42:837–846. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 66. Poorkaj P, Bird TD, Wijsman E, Nemens E, Garruto RM, Anderson L, Andreadis A, Widerholt WC, Raskind M, Schellenberg GD (1998) Tau is a candidate gene for chromosome 17 frontotemporal dementia. Ann Neurol 43:815–825. [DOI] [PubMed] [Google Scholar]
  • 67. Poorkaj P, Muma NA, Zhukareva V, Cochran EJ, Shannon KM, Hurtig H, Koller WC, Bird TD, Trojanowski JQ, Lee VM, Schellenberg GD (2002) An R5L tau mutation in a subject with a progressive supranuclear palsy phenotype. Ann Neurol 52: 511–516. [DOI] [PubMed] [Google Scholar]
  • 68. Rademakers R, Cruts M, Van Broeckhoven C (2004) The role of tau (MAPT) in frontotemporal dementia and related tauopathies. Hum Mutat 24:277–295. [DOI] [PubMed] [Google Scholar]
  • 69. Rademakers R, Melquist S, Cruts M, Theuns J, Del‐Favero J, Poorkaj P, Baker M, Sleegers K, Crook R, De Pooter T, Bel Kacem S, Adamson J, Van den Bossche D, Van den Broeck M, Gass J, Corsmit E, De Rijk P, Thomas N, Engelborghs S, Heckman M, Litvan I, Crook J, De Deyn PP, Dickson D, Schellenberg GD, Van Broeckhoven C, Hutton ML (2005) High‐density SNP haplotyping suggests altered regulation of tau gene expression in progressive supranuclear palsy. Hum Mol Genet 14:3281–3292. [DOI] [PubMed] [Google Scholar]
  • 70. Rajput A, Dickson DW, Robinson CA, Ross OA, Dachsel JC, Lincoln SJ, Cobb SA, Rajput ML, Farrer MJ (2006) Parkinsonism, Lrrk2 G2019S, and tau neuropathology. Neurology 67:1506–1508. [DOI] [PubMed] [Google Scholar]
  • 71. Rebeiz JJ, Kolodny EH, Richardson EP Jr (1968) Corticodentatonigral degeneration with neuronal achromasia. Arch Neurol 18:20–33. [DOI] [PubMed] [Google Scholar]
  • 72. Rojo A, Pernaute RS, Fontan A, Ruiz PG, Honnorat J, Lynch T, Chin S, Gonzalo I, Rabano A, Martinez A, Daniel S, Pramstaller P, Morris H, Wood N, Lees A, Tabernero C, Nyggard T, Jackson AC, Hanson A, De Yebenes JG (1999) Clinical genetics of familial progressive supranuclear palsy. Brain 122:1233–1245. [DOI] [PubMed] [Google Scholar]
  • 73. Ros R, Gomez Garre P, Hirano M, Tai YF, Ampuero I, Vidal L, Rojo A, Fontan A, Vazquez A, Fanjul S, Hernandez J, Cantarero S, Hoenicka J, Jones A, Ahsan RL, Pavese N, Piccini P, Brooks DJ, Perez‐Tur J, Nyggard T, De Yebenes JG (2005) Genetic linkage of autosomal dominant progressive supranuclear palsy to 1q31.1. Ann Neurol 57:634–641. [DOI] [PubMed] [Google Scholar]
  • 74. Ros R, Thobois S, Streichenberger N, Kopp N, Sanchez MP, Perez M, Hoenicka J, Avila J, Honnorat J, De Yebenes JG (2005) A new mutation of the tau gene, G303V, in early‐onset familial progressive supranuclear palsy. Arch Neurol 62:1444–1450. [DOI] [PubMed] [Google Scholar]
  • 75. Ross OA, Whittle AJ, Cobb SA, Hulihan MM, Lincoln SJ, Toft M, Farrer MJ, Dickson DW (2006) Lrrk2 R1441 substitution and progressive supranuclear palsy. Neuropathol Appl Neurobiol 32:23–25. [DOI] [PubMed] [Google Scholar]
  • 76. Rossi G, Gasparoli E, Pasquali C, Di Fede G, Testa D, Albanese A, Bracco F, Tagliavini F (2004) Progressive supranuclear palsy and Parkinson’s disease in a family with a new mutation in the tau gene. Ann Neurol 55:448. [DOI] [PubMed] [Google Scholar]
  • 77. Spillantini MG, Murrell JR, Goedert M, Farlow MR, Klug A, Ghetti B (1998) Mutation in the tau gene in familial multiple system tauopathy with presenile dementia. Proc Natl Acad Sci USA 95: 7737–7741. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 78. Spillantini MG, Yoshida H, Rizzini C, Lantos PL, Khan N, Rossor MN, Goedert M, Brown J (2000) A novel tau mutation (N296N) in familial dementia with swollen achromatic neurons and corticobasal inclusion bodies. Ann Neurol 48:939–943. [DOI] [PubMed] [Google Scholar]
  • 79. Stanford PM, Halliday GM, Brooks WS, Kwok JB, Storey CE, Creasey H, Morris JG, Fulham MJ, Schofield PR (2000) Progressive supranuclear palsy pathology caused by a novel silent mutation in exon 10 of the tau gene: expansion of the disease phenotype caused by tau gene mutations. Brain 123:880–893. [DOI] [PubMed] [Google Scholar]
  • 80. Steele JC, Richardson JC, Olszewski J (1964) Progressive Supranuclear Palsy. a Heterogeneous Degeneration Involving the Brain Stem, Basal Ganglia and Cerebellum with Vertical Gaze and Pseudobulbar Palsy, Nuchal Dystonia and Dementia. Arch Neurol 10:333–359. [DOI] [PubMed] [Google Scholar]
  • 81. Stefansson H, Helgason A, Thorleifsson G, Steinthorsdottir V, Masson G, Barnard J, Baker A, Jonasdottir A, Ingason A, Gudnadottir VG, Desnica N, Hicks A, Gylfason A, Gudbjartsson DF, Jonsdottir GM, Sainz J, Agnarsson K, Birgisdottir B, Ghosh S, Olafsdottir A, Cazier JB, Kristjansson K, Frigge ML, Thorgeirsson TE, Gulcher JR, Kong A, Stefansson K (2005) A common inversion under selection in Europeans. Nat Genet 37:129–137. [DOI] [PubMed] [Google Scholar]
  • 82. Tagliavini F, Pilleri G, Bouras C, Constantinidis J (1984) The basal nucleus of Meynert in patients with progressive supranuclear palsy. Neurosci Lett 44:37–42. [DOI] [PubMed] [Google Scholar]
  • 83. Tellez‐Nagel I, Wisniewski HM (1973) Ultrastructure of neurofibrillary tangles in Steele‐Richardson‐Olszewski syndrome. Arch Neurol 29: 324–327. [DOI] [PubMed] [Google Scholar]
  • 84. Togo T, Dickson DW (2002) Ballooned neurons in progressive supranuclear palsy are usually due to concurrent argyrophilic grain disease. Acta Neuropathol (Berl) 104:53–56. [DOI] [PubMed] [Google Scholar]
  • 85. Togo T, Cookson N, Dickson DW (2002) Argyrophilic grain disease: neuropathology, frequency in a dementia brain bank and lack of relationship with apolipoprotein E. Brain Pathol 12:45–52. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 86. Togo T, Sahara N, Yen SH, Cookson N, Ishizawa T, Hutton M, De Silva R, Lees A, Dickson DW (2002) Argyrophilic grain disease is a sporadic 4‐repeat tauopathy. J Neuropathol Exp Neurol 61:547–556. [DOI] [PubMed] [Google Scholar]
  • 87. Tolnay M, Probst A (1998) Ballooned neurons expressing alphaB‐crystallin as a constant feature of the amygdala in argyrophilic grain disease. Neurosci Lett 246:165–168. [DOI] [PubMed] [Google Scholar]
  • 88. Tsuboi Y, Ahlskog JE, Apaydin H, Parisi JE, Dickson DW (2001) Lewy bodies are not increased in progressive supranuclear palsy compared with normal controls. Neurology 57:1675–1678. [DOI] [PubMed] [Google Scholar]
  • 89. Tsuboi Y, Josephs KA, Cookson N, Dickson DW (2003) APOE E4 is a determinant for Alzheimer type pathology in progressive supranuclear palsy. Neurology 60:240–245. [DOI] [PubMed] [Google Scholar]
  • 90. Tsuboi Y, Slowinski J, Josephs KA, Honer WG, Wszolek ZK, Dickson DW (2003) Atrophy of superior cerebellar peduncle in progressive supranuclear palsy. Neurology 60:1766–1769. [DOI] [PubMed] [Google Scholar]
  • 91. Tsuboi Y, Josephs KA, Boeve BF, Litvan I, Caselli RJ, Caviness JN, Uitti RJ, Bott AD, Dickson DW (2005) Increased tau burden in the cortices of progressive supranuclear palsy presenting with corticobasal syndrome. Mov Disord 20:982–988. [DOI] [PubMed] [Google Scholar]
  • 92. Uchikado H, DelleDonne A, Ahmed Z, Dickson DW (2006) Lewy bodies in progressive supranuclear palsy represent an independent disease process. J Neuropathol Exp Neurol 65:387–395. [DOI] [PubMed] [Google Scholar]
  • 93. Urakami K, Wada K, Arai H, Sasaki H, Kanai M, Shoji M, Ishizu H, Kashihara K, Yamamoto M, Tsuchiya‐Ikemoto K, Morimatsu M, Takashima H, Nakagawa M, Kurokawa K, Maruyama H, Kaseda Y, Nakamura S, Hasegawa K, Oono H, Hikasa C, Ikeda K, Yamagata K, Wakutani Y, Takeshima T, Nakashima K (2001) Diagnostic significance of tau protein in cerebrospinal fluid from patients with corticobasal degeneration or progressive supranuclear palsy. J Neurol Sci 183:95–98. [DOI] [PubMed] [Google Scholar]
  • 94. Warren NM, Piggott MA, Perry EK, Burn DJ (2005) Cholinergic systems in progressive supranuclear palsy. Brain 128:239–249. [DOI] [PubMed] [Google Scholar]
  • 95. Williams DR, De Silva R, Paviour DC, Pittman A, Watt HC, Kilford L, Holton JL, Revesz T, Lees AJ (2005) Characteristics of two distinct clinical phenotypes in pathologically proven progressive supranuclear palsy: richardson’s syndrome and PSP‐parkinsonism. Brain 128:1247–1258. [DOI] [PubMed] [Google Scholar]
  • 96. Wszolek ZK, Pfeiffer RF, Tsuboi Y, Uitti RJ, McComb RD, Stoessl AJ, Strongosky AJ, Zimprich A, Muller‐Myhsok B, Farrer MJ, Gasser T, Calne DB, Dickson DW (2004) Autosomal dominant parkinsonism associated with variable synuclein and tau pathology. Neurology 62:1619–1622. [DOI] [PubMed] [Google Scholar]
  • 97. De Yebenes JG, Sarasa JL, Daniel SE, Lees AJ (1995) Familial progressive supranuclear palsy. Description of a pedigree and review of the literature. Brain 118:1095–1103. [DOI] [PubMed] [Google Scholar]
  • 98. Zimprich A, Biskup S, Leitner P, Lichtner P, Farrer M, Lincoln S, Kachergus J, Hulihan M, Uitti RJ, Calne DB, Stoessl AJ, Pfeiffer RF, Patenge N, Carbajal IC, Vieregge P, Asmus F, Muller‐Myhsok B, Dickson DW, Meitinger T, Strom TM, Wszolek ZK, Gasser T (2004) Mutations in LRRK2 cause autosomal‐dominant parkinsonism with pleomorphic pathology. Neuron 44:601–607. [DOI] [PubMed] [Google Scholar]

Articles from Brain Pathology are provided here courtesy of Wiley

RESOURCES