ArticlePDF Available

Patients with vascular damage due to microvascular pathology have significant hippocampal neuronal loss

Authors:

Abstract and Figures

Alzheimer's disease (AD) is characterised by functional impairment, cerebral atrophy, and degeneration of specific neuronal populations, especially pyramidal neurones of the cerebral cortex and hippocampal formation. Although patients with subcortical vascular dementia have been shown to have similar metabolic and volumetric deficits to those with AD, the underlying pathogenesis of these changes is poorly understood. To determine whether pyramidal cell loss occurs in small vessel disease (SVD) dementia by quantifying hippocampal volume and CA1 neurone number. Fifty four prospectively studied patients with dementia were screened, and four patients fulfilling criteria for SVD with no other significant neuropathological abnormality were identified. These were compared with five patients fulfilling criteria for AD and seven controls matched for age and sex. The hippocampal formation was serially sectioned, and the number of CA1 pyramidal neurones estimated using the optical dissector technique. Analysis of variance was used to evaluate group differences. Patients in both the AD and SVD groups showed a substantial loss of pyramidal neurones from the CA1 region. The pattern of hippocampal atrophy and the degree of CA1 neuronal loss were similar in the two dementia groups. These findings support recent in vivo studies showing similar metabolic deficits and atrophy in AD and subcortical vascular dementia. In addition, they provide evidence that the underlying cause of these abnormalities is a similar loss of neurones. Whereas the cause of the neuronal loss in AD is related to the deposition of abnormal proteins, the cause in SVD is unknown. In the absence of other pathologies, damage to cerebral microvasculature should be considered a likely candidate.
Content may be subject to copyright.
PAPER
Patients with vascular dementia due to microvascular
pathology have significant hippocampal neuronal loss
J J Kril, S Patel, A J Harding, G M Halliday
.............................................................................................................................
J Neurol Neurosurg Psychiatry
2002;72:747–751
Background: Alzheimer’s disease (AD) is characterised by functional impairment, cerebral atrophy,
and degeneration of specific neuronal populations, especially pyramidal neurones of the cerebral cor-
tex and hippocampal formation. Although patients with subcortical vascular dementia have been
shown to have similar metabolic and volumetric deficits to those with AD, the underlying pathogenesis
of these changes is poorly understood.
Objective: To determine whether pyramidal cell loss occurs in small vessel disease (SVD) dementia by
quantifying hippocampal volume and CA1 neurone number.
Methods: Fifty four prospectively studied patients with dementia were screened, and four patients ful-
filling criteria for SVD with no other significant neuropathological abnormality were identified. These
were compared with five patients fulfilling criteria for AD and seven controls matched for age and sex.
The hippocampal formation was serially sectioned, and the number of CA1 pyramidal neurones esti-
mated using the optical dissector technique. Analysis of variance was used to evaluate group
differences.
Results: Patients in both the AD and SVD groups showed a substantial loss of pyramidal neurones from
the CA1 region. The pattern of hippocampal atrophy and the degree of CA1 neuronal loss were simi-
lar in the two dementia groups.
Conclusions: These findings support recent in vivo studies showing similar metabolic deficits and atro-
phy in AD and subcortical vascular dementia. In addition, they provide evidence that the underlying
cause of these abnormalities is a similar loss of neurones. Whereas the cause of the neuronal loss in
AD is related to the deposition of abnormal proteins, the cause in SVD is unknown. In the absence of
other pathologies, damage to cerebral microvasculature should be considered a likely candidate.
Vascular dementia is recognised as one of the most com-
mon causes of dementia after Alzheimer’s disease (AD).
Traditionally, vascular dementia has been distinguished
from AD by the pattern of clinical progression, with vascular
dementia characterised by a stepwise rather than a slowly
progressive course of deterioration. However, this view
appears overly simplistic, as clinical presentations can vary
enormously in patients with vascular dementia.1–5 In particu-
lar, some cases of vascular dementia are clinically indistin-
guishable from AD as they show no evidence of a stepwise
deterioration.2At autopsy, these patients have small vessel
disease (SVD), leucoencephalopathy, and microscopic infarc-
tion rather than the pathology of AD.2This has been confirmed
in a carefully selected autopsy series, leading Esiri and
colleagues4to suggest that microvascular disease, rather than
macroscopic infarction, is the most common pathology
underlying vascular dementia.
Although SVD is being increasingly recognised as a subtype
of vascular dementia,56no validated criteria for the identifica-
tion of such cases, either clinically or pathologically, have been
established. At present, the pathological classification of SVD
is largely descriptive and often relies on the pathologist’s
opinion of whether the vascular pathology contributed to the
dementia.7The study by Esiri and colleagues4proposed a
semiquantitative grading scale for SVD using the presence and
increasing severity of a number of features of microvascular
pathology.The systematic use of this scale may lead to a better
understanding of the mechanisms by which SVD causes, or
contributes to, dementia.
An increase in SVD is known to occur with age and in the
presence of cerebrovascular risk factors.8In non-demented
persons, microvascular pathology influences specific cognitive
functions such as those involving speed of mental
processing.8Although the substrate for dementia in SVD is
still unclear, in vivo studies910indicate a similar pattern and
degree of cortical dysfunction and volume loss in AD and SVD.
The functional deficits reported to date in patients with SVD
dementia suggest significant cortical and limbic degeneration,
similar to AD, although the underlying pathology is predomi-
nantly white matter leucoencephalopathy. Remote hippocam-
pal injury has been found in some cases with extensive micro-
vascular disease,11 but has not been systematically studied.
Indeed, there have been no studies on the distribution and
extent of neuronal loss in patients with SVD. In this study,
neuronal loss from the CA1 sector of the hippocampus was
quantified in demented patients who did not reach pathologi-
cal criteria for AD, or any other neurodegenerative disease,but
who had significant SVD. The aim of this study was to deter-
mine if SVD per se can cause hippocampal degeneration and
as such contribute to the dementia syndrome.
MATERIALS AND METHODS
Case selection and characterisation
Cases were collected from both population based research
studies on AD and from a tertiary referral centre for neuro-
degenerative diseases. This study was approved by the ethics
review committees of the Central and South Eastern Sydney
Area Health Services. All patients were examined by a
neurologist or geriatrician. Additional corroborative infor-
mation for each patient was obtained from an informant
interview to ascertain the pattern and type of deficits and to
.............................................................
Abbreviations: AD, Alzheimer’s disease; SVD, small vessel disease
See end of article for
authors’ affiliations
.......................
Correspondence to:
Dr J J Kril, Centre for
Education and Research on
Ageing, Concord Hospital,
Concord NSW 2139,
Australia;
jilliank@med.usyd.edu.au
Received
12 November 2001
Accepted
15 January 2002
.......................
747
www.jnnp.com
aid in the estimation of disease duration. Fifty four patients
who met NINCDS-ADRDA criteria for possible or probable
AD12 came to autopsy during the collection period of seven
years.
Brains were fixed in neutral buffered formalin for two
weeks, and then the weight, volume, and anteroposterior
length of the cerebral hemispheres determined.13 The hemi-
spheres were then sectioned at about 3 mm intervals in the
coronal plane. The cerebral cortex, hippocampus, brainstem,
and cerebellum were sampled from standardised regions of
each brain. Sections were stained with haematoxylin and
eosin, modified Bielschowsky silver and ubiquitin immuno-
histochemistry, and examined to perform the classifications
outlined below and to exclude pathologies other than AD and
SVD.
Each case was given a grading using each of the following
classification schema.
(a) CERAD14: normal, possible, probable, or definite AD.
Neuritic plaques were evaluated in sections of the temporal
neocortex (Brodmann area 20) stained with the modified
Bielschowsky silver stain. Ten random microscope fields at
100×magnification graded as having none (<1/field), rare
(2–5/field), mild (6–15/field), moderate (16–50/field), or
severe (>50/field) numbers of neuritic plaques.
(b) Braak15 16: normal (stages 0–2), transitional limbic (stages
3 or 4), or neocortical (stages 5 or 6) AD. Neurofibrillary tan-
gles were evaluated in sections of the hippocampal formation
and temporal neocortex stained with the modified Biels-
chowsky silver stain. Ten random microscope fields at 200×
magnification graded into none (nil found), rare (<1/field),
mild (1–2/field), moderate (3–10/field) or severe (>10/field)
numbers of neurofibrillary tangles.
(c) Esiri4: normal (grades 0 or 1), SVD (grades 2 or 3). Sections
from the dorsolateral prefrontal association cortex (Brod-
mann area 9), anterior cingulate gyrus (Brodmann area 24),
and primary motor cortex (Brodmann area 4) were stained
with haematoxylin and eosin. Briefly, cases were classified as
normal if they exhibited little or no widening of perivascular
spaces, thickening of vascular walls, and/or a few perivascular
macrophages. Notably, myelin pallor, attenuation of nerve
fibres, and gliosis were absent. Cases were classified as SVD
when widening of perivascular spaces (fig 1C), thickening of
vascular walls (fig 1A), accumulation of perivascular macro-
phages (fig 1B), and perivascular myelin pallor or attenuation
of nerve fibres (fig 1D) were present.
For this study, case classification for SVD was based on the
presence of SVD using the Esiri grading scheme and the
absence of probable or definite CERAD AD or a Braak neocor-
tical neuritic stage of AD. Four of the 54 cases fulfilled such
criteria for SVD in the absence of any other significant degen-
erative condition. For comparison, five AD cases were selected
on the basis of absence of SVD and the presence of definite
CERAD AD and a Braak neocortical stage of AD (table 1).
Although 12 of the 54 cases showed both SVD and AD using
the classifications outlined above (four with additional large
vessel infarction, two with dementia with Lewy bodies,17 and
two with hippocampal sclerosis), these cases were not
included in the study. This enabled examination of the role of
SVD, and its comparison with AD, without the complication of
an overlap in the type of cortical pathology in the two demen-
tia groups. Seven controls matched for age and sex from the
population based research studies who were without signifi-
cant neuropathology (normal using all schema) were also
selected for comparison.
Quantification of hippocampal formation
The techniques used to quantify hippocampal volume and
CA1 neurone number have been described in detail in our
previous publications.18 19 Briefly, the entire hippocampal
formation was dissected from the right hemisphere, the blocks
cryoprotected in 30% sucrose, and three 48 µm thick sections
cut from the caudal face of each block on a freezing
microtome. Series of sections from each case were stained
Figure 1 Photomicrographs showing white matter pathology used in the grading of small vessel disease pathology. (A) Haematoxylin and
eosin stained section through the frontal white matter showing an arteriole with a thickened vessel wall. Scale equivalent to (B). (B)
Haematoxylin and eosin stained section through the frontal white matter showing haemosiderin laden macrophages around a white matter
blood vessel. (C) Haematoxylin and eosin stained section through the frontal white matter showing perivascular dilatation with attenuation of
surrounding white matter. (D) Modified Bielschowsky silver stained section through the frontal white matter showing attenuation of myelin in the
deep white matter.
748 Kril, Patel, Harding, et al
www.jnnp.com
with haematoxylin and eosin, cresyl violet, and nickel
peroxidase.20 The boundaries of each of the hippocampal sub-
regions were delineated on the cresyl violet stained sections
and drawn after magnification on a microfiche reader (×19
magnification). The volume of each region was then
determined using point counting,18 19 and the total number of
neurones in the CA1 region estimated using the optical
dissector technique.18 19 21 22 The first dissector frame (120 µm ×
120 µm) was placed randomly in the CA1 sector,and then sys-
tematic samples at 2.4 mm horizontally and 1.2 mm vertically
were counted. The coefficient of variance for the control group
was 0.208, and the coefficient of error was 0.074.
Group means were compared using analysis of variance,
and correlations examined using linear regression analysis
(Statview 5, SAS Institute Inc, Cary, North Carolina, USA).
RESULTS
Case descriptions
Table 2 gives the details of the classification for each case. The
mean age did not differ between any of the groups (F=2.61;
p=0.112), although patients in the SVD group were on
average older than those in the AD groups (86 (12) years v76
(4) years; controls had an mean age of 75 (8) years) (values
are mean (SD)).
Summaries of the clinical presentation and progression of
each case with SVD are listed below to illustrate more clearly
the type of case studied.
S1 was a 69 year old woman who presented with increasing
impairment of memory. Her past medical history consisted of
type II diabetes, hyperlipidaemia, and hypertension, which
was controlled by captopril. After evaluation she was
diagnosed with AD. The following year she suffered a stroke-
like episode, but computed tomography showed no focal
lesion. Cerebral atrophy and leucoaraiosis were described. She
continued to decline and was cared for at home by her
husband. Three years after presentation she was admitted to a
regional hospital after a fall. She died nine days later after a
five day history of decline in neurological function. Death was
attributed to an acute myocardial infarct. Her neuropathologi-
cal examination showeda3mmlacune in the right putamen
and an area of destruction of the white matter at the angle of
the right lateral ventricle.
S2 was a university professor who retired at 65 years of age,
although had continued to participate in academic activities.
He presented at age 75 with a two year history of increasing
difficulty with word finding and had ceased lecturing because
of this. He had also developed difficulties with navigation
while driving. Formal assessment at this time revealed a mini
mental state examination23 of 30/35, and a diagnosis of AD was
made. He had no history of diabetes or hypertension, although
he had suffered a transient ischaemic attack two years earlier.
His medications were aspirin, melleril, and voltaren. Over the
ensuing three years he deteriorated, became inert, and was
unable to handle money, although he was able to dress and
groom himself. Neurological examination showed brady-
kinesia, rigidity, and ataxia. He died at age 80, and
neuropathological examination showed small lacunes in the
left internal capsule, right claustrum, and right globus
pallidus. There was some loss of pigmented neurones from the
substantia nigra but no Lewy bodies.
S3 was a 93 year old woman referred for evaluation. She had
a 10 year history of a decline in domestic duties and increas-
ing neglect of her self care. She had delusions and wandered
if left unattended. On examination, she had deficits in
memory, insight, visuospatial abilities, drive, and planning.
Her mini mental state examination was 14/30. Her previous
history included a myocardial infarct eight years earlier, but
was otherwise unremarkable. She had several admissions to
hospital after falls, including a fractured pelvis. She was
admitted to a nursing home one year before death.
Neuropathological examination showed an old cavitated
infarct involving the internal capsule, caudate, and putamen,
which measured 1.5 cm in anterioposterior extent. Haemo-
siderin laden macrophages were present at this site.
S4 was a 94 year old woman who presented with a one year
history of increasing forgetfulness. She had a sustained delu-
sion that her symptoms were due to her use of a pesticide
spray. She also had visual hallucinations. Investigation
Table 1 Screening procedure and classification
schema
Diagnosis CERAD14 Braak16 Esiri4
SVD alone Normal or possible Normal (0–2) SVD (2–3)
or limbic (3–4)
AD alone Probable or definite Neocortical (5–6) Normal (0–1)
All cases met NINCDS-ADRDA34 clinical criteria for possible or
probable AD. Note the lack of overlap between the two groups. The
combination of the CERAD and Braak schemes for AD is consistent
with current recommendations for the diagnosis of AD.35
SVD, Small vessel disease; AD, Alzheimer’s disease.
Table 2 Age, sex, and neuropathological grading for each case
Case
Age
(years) Sex
Duration
(years)
Plaque
density CERAD*
NFT density
in ctx Braak*
SVD*
grade
S1 73 F 4 6–15 Possible 0 Normal 3
S2 80 M 7 Rare Normal Rare Limbic 3
S3 96 F 13 Rare Normal 1–2 Limbic 3
S4 97 F 4 6–15 Possible Rare Limbic 2
A5 69 F 4 16–50 Definite >10 Neocortical 1
A6 70 F 7 >50 Definite >10 Neocortical 1
A7 76 M 4 >50 Definite >10 Neocortical 0
A8 78 M 14 16–50 Definite >10 Neocortical 1
A9 81 F 5 16–50 Definite 3–10 Neocortical 0
C1 61 F Rare Normal 0 Normal 1
C2 69 F Rare Normal 0 Normal 0
C3 72 M Rare Normal 0 Normal 0
C4 74 M Rare Normal 0 Normal 0
C5 79 M Rare Normal 0 Normal 1
C6 84 F Rare Normal 0 Normal 1
C7 85 M Rare Normal 0 Normal 1
*CERAD14; Braak16; Esiri.4
Ctx, Cortex; NFT, neurofibrillary tangles; S, SVD dementia; A, Alzheimer’s disease; C, non-demented control.
Hippocampal neuronal loss in vascular dementia 749
www.jnnp.com
showed deficits in memory, praxis, and an unsteady gait con-
sistent with having had bilateral knee replacements. She had
no history of diabetes, cardiac disease, or hypertension and
was taking no medicines. She died from a cardiac arrest while
resident in a nursing home. Neuropathological examination
showed generalised atrophy, with no focal lesions.
Diagnostic pathology
Figure 1 shows examples of the white matter pathology in the
SVD cases. Changes included both those involving the vessels
themselves and those involving the surrounding tissue. Thick-
ening of arteriolar walls (fig 1A), accumulation of macro-
phages (fig 1B), perivascular dilatation (etat criblé, fig 1C),
and attenuation of myelin (fig 1D) were seen in all cases.
These changes were observed in all of the white matter regions
examined.
In the SVD cases, no cortical plaques were detected in two
cases and moderate numbers were found in the other two
(table 2). Cortical neurofibrillary tangles were absent in three
cases and mild in the fourth (table 2). In contrast, the patients
with AD had severe cortical neurofibrillary tangles in four of
the five cases and moderate in the fifth (table 2). Cortical
plaques were severe in two of the five and moderate in the
remainder (table 2).
Hippocampal volume changes
The grey matter component of the hippocampal formation is
significantly smaller in both of the dementia groups than in
the control group (table 3). When the subregion volumes were
analysed, all regions except the dentate gyrus and hilus (CA4)
were smaller in patients with AD than in controls. In SVD, sig-
nificant atrophy was confined to the large CA1 and subiculum
grey matter regions (table 3). The magnitude of the atrophy in
these hippocampal regions was not significantly different
between the dementia groups.
CA1 neurone number
The total estimated neurone number in the CA1 is also
significantly reduced in both dementia groups compared with
controls (F=17.3; p=0.0002; fig 2).The magnitude of CA1 loss
is not significantly different between the SVD and AD groups
(50% and 69% respectively). Considerable variation in
neurone number exists within the two dementia groups,
although this is not as a result of disease duration (r2=0.035;
p=0.63) or age at death (r2=0.001; p=0.94)
DISCUSSION
Significant atrophy and neuronal loss from the CA1 region of
the hippocampus are one of the pathological hallmarks of
AD.22 24 Interestingly in this study we have shown that a simi-
lar pattern and magnitude of hippocampal atrophy and
neurone loss also occurs in patients who do not reach patho-
logical criteria for AD but who have SVD. These findings,
together with the fact that all nine patients with dementia
met clinical criteria for probable AD, suggest that SVD can
result in a cortical dementia syndrome similar to AD and that
hippocampal neuronal loss contributes to this clinical picture.
Hippocampal damage in AD is considered an early event
with substantial pathology and early atrophy preceding symp-
tom onset.25 26 After these early changes, the widespread depo-
sition of abnormal proteins within the association cortices
occurs, with cognitive decline leading to clinical dementia and
appreciable neurodegeneration.16 27 28 This knowledge of the
progression of pathology underlying the clinical dementia and
cortical degeneration of AD has been derived from studying a
large number of cases. Similar studies have not been
performed for SVD dementia because of the lack of consensus
on diagnostic requirements for SVD and the relative rareness
of cases in which SVD is the only substrate for the dementia.
In our prospectively studied series of patients with dementia,
and other autopsy series,24 the proportion of cases in which
significant microvascular disease is the principal pathology is
small (<10% in this study). Several studies have investigated
the consequences of coexisting cerebrovascular disease and
AD. Both the OPTIMA and Nun studies show that cognitive
performance is worse in subjects with early AD and cerebro-
vascular disease (both large vessel and small vessel disease)
than in patients with AD alone,29 30 suggesting that the vascu-
lar pathology contributes to dementia severity. Similarly, in
subjects with equivalent cognitive performance, the burden of
AD-type pathology is less in those with AD and coexisting
cerebrovascular disease than in those with AD alone,731 also
supporting the concept that vascular pathology contributes to
the dementia. Our results suggest that this contribution is due
to pyramidal cell loss through non-AD mechanisms.
The study of vascular dementia has been dominated by
large vessel disease (infarction) and only recently has the
importance of SVD been realised.411The loss of neurones with
large vessel disease is unquestioned, although neuronal loss
due to microvascular pathology has been anecdotal and not
systematically studied. Our study not only provides evidence
for the existence of SVD as an independent cause of dementia,
but also provides some insights into the pathogenesis of this
dementia by showing that substantial hippocampal pyramidal
cell loss occurs in SVD. This suggests a causative link between
Table 3 Volumes (mm3) of the hippocampal subregions for each group
SVD p Value AD p Value Control
Total 1302 (142) 0.0053 1035 (74) 0.0001 1835 (108)
CA1 466 (77) 0.0270 317 (46) 0.0004 662 (46)
CA2–3 119 (10) 0.5813 89 (14) 0.0344 129 (12)
CA4 160 (8) 0.4187 158 (12) 0.3516 185 (25)
Dentate gyrus 51 (2) 0.6050 59 (5) 0.0924 47 (4)
Subiculum 254 (37) 0.0112 211 (23) 0.0025 490 (66)
Presubiculum 251 (36) 0.0668 201 (26) 0.0028 322 (17)
Values are mean (SEM). Values significantly different from controls are shown in bold.
SVD, Small vessel disease; AD, Alzheimer’s disease.
Figure 2 Neurone number in the CA1 region of the hippocampus
in control, Alzheimer’s disease (AD) and small vessel disease (SVD)
groups. Bar = mean ±1 SD, p values for AD and SVD groups
relative to control group.
8
7
6
5
4
3
2
1
0Control AD SVD
p = 0.0029
p < 0.0001
10–6 × Total CA1 neurone number
750 Kril, Patel, Harding, et al
www.jnnp.com
SVD, pyramidal cell loss, and hippocampal degeneration, and
raises the possibility that cortical microvascular pathology
contributes to dementia through pyramidal cell loss. In vivo
studies showing a similar pattern and severity of cortical dys-
function and atrophy in AD and SVD support this
suggestion.910 Overall, there may be little difference in the
degree of neuronal loss between SVD and AD, with only the
mechanism of degeneration differing between these dementia
syndromes.
The mechanism of neuronal degeneration in SVD may be
apoptosis, consistent with an absence of pathological debris.
Apoptosis is increased in ischaemia,32 and markers of apopto-
sis are increased in areas of leucoaraiosis compared with adja-
cent white matter.33 An increase in cell cycle proteins within
hippocampal neurones in patients with SVD31 is also
consistent with abnormal cellular processing leading to apop-
tosis. The location of these changes within hippocampal CA1
neurones suggests they may precede the cell loss shown in this
study. It is now important to study large cohorts of patients
with SVD in order to determine the temporal sequence of the
neuronal damage shown and the onset of dementia. Such
knowledge is essential for effective treatment strategies.
ACKNOWLEDGEMENTS
This work was supported by a programme grant from the Medical
Foundation of the University of Sydney and by the National Health
and Medical Research Council of Australia. JJK was a Medical Foun-
dation Fellow, and GMH is a NHMRC Principal Research Fellow.
.....................
Authors’ affiliations
JJKril,SPatel,Centre for Education and Research on Ageing,
Department of Medicine, University of Sydney, Sydney, NSW, Australia
J J Kril, Department of Pathology, University of Sydney
A J Harding, G M Halliday, Prince of Wales Medical Research Institute
and the University of New South Wales, Australia
REFERENCES
1Wallin A. The overlap between Alzheimer’s disease and vascular
dementia: the role of white matter changes.
Dement Geriatr Cogn Disord
1998;9(suppl 1):30–5.
2Pantoni L, Garcia JH, Brown GG. Vascular pathology in three cases of
progressive cognitive decline.
J Neurol Sci
1996;135:131–9.
3Hachinski V, Norris JW. Vascular dementia: an obsolete concept.
Curr
Opin Neurol
1994;7:3–4.
4Esiri MM, Wilcock GK, Morris JH. Neuropathological assessment of the
lesions of significance in vascular dementia.
J Neurol Neurosurg
Psychiatry
1997;63:749–53.
5Englund E. White matter pathology of vascular dementia. In: Chiu E,
Gustafson L, Ames D,
et al
,eds.
Cerebrovascular disease and dementia.
Pathology, neuropsychiatry and management.
London: Martin Dunitz Ltd,
2000:77–84.
6Erkinjuntti T, Benavente O, Eliasziw M,
et al
. Diffuse vacuolization
(spongiosis) and arteriolosclerosis in the frontal white matter occurs in
vascular dementia.
Arch Neurol
1996;53:325–32.
7Nagy Z, Esiri MM, Jobst KA,
et al
. The effects of additional pathology on
the cognitive deficit in Alzheimer disease.
J Neuropathol Exp Neurol
1997;56:165–70.
8Pantoni L, Garcia JH. The significance of cerebral white matter
abnormalities 100 years after Binswager’s report: a review.
Stroke
1995;26:1293–301.
9Mielke R, Heiss W-D. Positron emission tomography for diagnosis of
Alzheimer’s disease and vascular dementia.
J Neural Trans
1998;53:237–50.
10 Pantel J, Schroder J, Essig M,
et al
. In vivo quantification of brain
volumes in subcortical vascular dementia and Alzheimer’s disease. An
MRI-based study.
Dement Geriatr Cogn Disord
1998;9:309–16.
11 Vinters HV, Ellis WG, Zarow C,
et al
. Neuropathologic substrates of
ischemic vascular dementia.
J Neuropathol Exp Neurol
2000;59:931–45.
12 Morris JC, Heyman A, Mohs RC,
et al
. The consortium to establish a
registry for Alzheimer’s disease (CERAD). Part I. Clinical and
neuropsychological assessment of Alzheimer’s disease.
Neurology
1989;39:1159–65.
13 Kril JJ, Halliday GM, Svoboda MD,
et al
. The cerebral cortex is
damaged in chronic alcoholics.
Neuroscience
1997;79:983–98.
14 Mirra SS, Heyman A, McKeel D,
et al
. The consortium to establish a
registry for Alzheimer’s disease (CERAD). Part II. Standardisation of the
neuropathologic assessment of Alzheimer’s disease.
Neurology
1991;41:479–86.
15 Harding AJ, Kril JJ, Halliday GM. A simplified Braak tangle staging
method for routine pathological screening.
Acta Neuropathol
2000;99:199–208.
16 Braak H, Braak E. Neuropathological stageing of Alzheimer-related
changes.
Acta Neuropathol
1991;82:239–59.
17 Harding AJ, Halliday GM. Simplified neuropathological diagnosis of
dementia with Lewy bodies.
Neuropathol Appl Neurobiol
1998;24:195–201.
18 Harding AJ, Halliday GM, Kril JJ. Variation in hippocampal neuron
number with age and brain volume.
Cereb Cortex
1998;8:710–18.
19 Harding AJ, Wong A, Svoboda M,
et al
. Chronic alcohol consumption
does not cause hippocampal neuron loss in humans.
Hippocampus
1997;7:1–10.
20 Cullen K. A novel nickel peroxidase method for the histochemical
visualization of neurofibrillary tangles, senile plaques and neuropil
threads.
J Histochem Cytochem
1994;42:1383–91.
21 West MJ. Regionally specific loss of neurons in the aging human
hippocumpus.
Neurobiol Aging
1993;14:287–93.
22 West MJ, Coleman PD, Flood DG,
et al
. Differences in the pattern of
hippocampal neuronal loss in normal ageing and Alzheimer’s disease.
Lancet
1994;344:769–72.
23 Folstein MF, Folstein SE, McHugh PR. “Mini-mental state”: a practical
method for grading the cognitive state of patients for the clinician.
J
Psychiatr Res
1975;12:189–98.
24 Ball MJ. Neuronal loss, neurofibrillary tangles and granulovacuolar
degeneration in the hippocampus with ageing and dementia: a
quantitative study.
Acta Neuropathol
1977;37:111–18.
25 Smith AD, Jobst KA. Use of structural imaging to study the progression of
Alzheimer’s disease.
Br Med Bull
1996;52:575–86.
26 Smith AD, Jobst KA, Edmonds Z,
et al
. Neuroimaging and early
Alzheimer’s disease.
Lancet
1996;348:829–30.
27 Gomez-Isla T, Hollister R, West H,
et al
. Neuronal loss correlates with
but exceeds neurofibrillary tangles in Alzheimer’s disease.
Ann Neurol
1997;41:17–24.
28 Convit A, de Asis J, de Leon MJ,
et al
. Atrophy of the medial
occipitotemporal, inferior, and middle temporal gyri in non-demented
elderly predict decline to Alzheimer’s disease.
Neurobiol Aging
2000;21:19–26.
29 Esiri MM, Nagy Z, Smith MZ,
et al
. Cerebrovascular disease and
threshold for dementia in the early stages of Alzheimer’s disease.
Lancet
1999;354:919–20.
30 Snowdon DA, Greiner LH, Mortimer JA,
et al
. Brain infarction and the
clinical expression of Alzheimer’s disease. The Nun Study.
JAMA
1997;277:813–17.
31 Smith MZ, Nagy Z, Barnetson L,
et al
. Coexisting pathologies in the
brain: influence of vascular disease and Parkinson’s disease on
Alzheimer’s pathology in the hippocampus.
Acta Neuropathol
2000;100:87–94.
32 Snider BJ, Gottron FJ, Choi DW. Apoptosis and necrosis in
cerebrovascular disease.
Ann N Y Acad Sci
1999;893:243–53.
33 Brown WR, Moody DM, Thore CR,
et al
. Apoptosis in leukoaraiosis.
AJNR Am J Neuroradiol
2000;21:79–82.
34 Tierney MC, Fisher RH, Lewis AJ,
et al
. The NINCDS-ADRDA work
group criteria for the clinical diagnosis of probable Alzheimer’s disease:
a clinicopathological study of 57 cases.
Neurology
1988;38:359–64.
35 National Institute on Aging and Reagan Institute Working Group on
diagnostic criteria for the neuropathological diagnosis of Alzheimer’s
disease. Consensus recommendations for the postmortem diagnosis of
Alzheimer’s disease.
Neurobiol Aging
1997;18(4S):S1–3.
Hippocampal neuronal loss in vascular dementia 751
www.jnnp.com
... We found hyperglycemia was an independent risk factor for hippocampal atrophy, which was similar to other studies (30,31). Additionally, we also found combined cerebral microvascular diseases was another independent risk factor for hippocampal atrophy (32). ...
Article
Full-text available
Background Cognitive impairment and brain atrophy are common in chronic kidney disease patients. It remains unclear whether differences in renal function, even within normal levels, influence hippocampal volume (HCV) and cognition. We aimed to investigate the association between estimated glomerular filtration rate (eGFR), HCV and cognition in outpatients. Methods This single-center retrospective study enrolled 544 nonrenal outpatients from our hospital. All participants underwent renal function assessment and 3.0 T magnetic resonance imaging (MRI) in the same year. HCV was also measured, and cognitive assessments were obtained. The correlations between eGFR, HCV, and cognitive function were analyzed. Logistic regression analysis was performed to identify the risk factors for hippocampal atrophy and cognitive impairment. Receiver-operator curves (ROCs) were performed to find the cut-off value of HCV that predicts cognitive impairment. Results The mean age of all participants was 66.5 ± 10.9 years. The mean eGFR of all participants was 88.5 ± 15.1 mL/min/1.73 m². eGFR was positively correlated with HCV and with Mini-Mental State Examination (MMSE) and Montreal Cognitive Assessment (MoCA) scores. Univariate and multivariate logistic regression analysis showed Age ≥ 65 years, eGFR < 75 mL/min/1.73 m², Glucose ≥6.1 mmol/L and combined cerebral microvascular diseases were independent risk factors for hippocampal atrophy and Age ≥ 65 years, left hippocampal volume (LHCV) <2,654 mm³ were independent risk factors for cognitive impairment in outpatients. Although initial unadjusted logistic regression analysis indicated that a lower eGFR (eGFR < 75 mL/min/1.73 m²) was associated with poorer cognitive function, this association was lost after adjusting for confounding variables. ROC curve analysis demonstrated that LHCV <2,654 mm³ had the highest AUROC [(0.842, 95% CI: 0.808–0.871)], indicating that LHCV had a credible prognostic value with a high sensitivity and specificity for predicting cognitive impairment compared with age in outpatients. Conclusion Higher eGFR was associated with higher HCV and better cognitive function. eGFR < 75 mL/min/1.73 m² was an independent risk factor for hippocampal atrophy after adjusting for age. It is suggested that even eGFR < 75 mL/min/1.73 m², lower eGFR may still be associated with hippocampal atrophy, which is further associated with cognitive impairment. LHCV was a favorable prognostic marker for predicting cognitive impairment rather than age.
... Meanwhile, research has depicted that the CA1 subfield of the hippocampus in gerbils appears to be vulnerable to anoxic-ischemic insults (Huang et al., 2022;Kirino & Sano, 1984). Further, autopsy studies of patients with microvascular pathology have demonstrated notable atrophy and neuronal loss in the CA1 region (Kril et al., 2002). Consequently, alterations identified in neurodegenerative and vascular cognitive impairments may imply a common pathological mechanism for cognitive decline. ...
Article
Full-text available
Early diagnosis of subcortical vascular mild cognitive impairment (svMCI) is clinically essential because it is the most reversible subtype of all cognitive impairments. Since structural alterations of hippocampal sub-regions have been well studied in neurodegenerative diseases with pathophysiological cognitive impairments, we were eager to determine whether there is a selective vulnerability of hippocampal sub-fields in patients with svMCI. Our study included 34 svMCI patients and 34 normal controls (NCs), with analysis of T1 images and Montreal Cognitive Assessment (MoCA) scores. Gray matter volume (GMV) of hippocampal sub-regions was quantified and compared between the groups, adjusting for age, sex, and education. Additionally, we explored correlations between altered GMV in hippocampal sub-fields and MoCA scores in svMCI patients. Patients with svMCI exhibited selectively reduced GMV in several left hippocampal sub-regions, such as the hippocampal tail, hippocampal fissure, CA1 head, ML-HP head, CA4 head, and CA3 head, as well as decreased GMV in the right hippocampal tail. Specifically, GMV in the left CA3 head was inversely correlated with MoCA scores in svMCI patients. Our findings indicate that the atrophy pattern of patients with svMCI was predominantly located in the left hippocampal sub-regions. The left CA3 might be a crucial area underlying the distinct pathophysiological mechanisms of cognitive impairments with subcortical vascular origins.
... Notably, the hippocampus is an important component of the temporal lobe, consisting of several subfields such as CA, subiculum, presubiculum, parasubiculum, dentate gyrus, and molecular layer (Brown et al., 2020). An autopsy study found reduced volume and neuronal loss in the hippocampus in SIVD patients with dementia (Kril et al., 2002). Li et al. (2016) further observed that patients with mild subcortical vascular cognitive impairment exhibited significant atrophy in the subiculum, presubiculum and dentate gyrus. ...
Article
Full-text available
Introduction Prior MRI studies have shown that patients with subcortical ischemic vascular disease (SIVD) exhibited white matter damage, gray matter atrophy and memory impairment, but the specific characteristics and interrelationships of these abnormal changes have not been fully elucidated. Materials and methods We collected the MRI data and memory scores from 29 SIVD patients with cognitive impairment (SIVD-CI), 29 SIVD patients with cognitive unimpaired (SIVD-CU) and 32 normal controls (NC). Subsequently, the thicknesses and volumes of the gray matter regions that are closely related to memory function were automatically assessed using FreeSurfer software. Then, the volume, fractional anisotropy (FA), mean diffusivity (MD), amplitude of low-frequency fluctuation (ALFF) and regional homogeneity (ReHo) values of white matter hyperintensity (WMH) region and normal-appearing white matter (NAWM) were obtained using SPM, DPARSF, and FSL software. Finally, the analysis of covariance, spearman correlation and mediation analysis were used to analyze data. Results Compared with NC group, patients in SIVD-CI and SIVD-CU groups showed significantly abnormal volume, FA, MD, ALFF, and ReHo values of WMH region and NAWM, as well as significantly decreased volume and thickness values of gray matter regions, mainly including thalamus, middle temporal gyrus and hippocampal subfields such as cornu ammonis (CA) 1. These abnormal changes were significantly correlated with decreased visual, auditory and working memory scores. Compared with the SIVD-CU group, the significant reductions of the left CA2/3, right amygdala, right parasubiculum and NAWM volumes and the significant increases of the MD values in the WMH region and NAWM were found in the SIVD-CI group. And the increased MD values were significantly related to working memory scores. Moreover, the decreased CA1 and thalamus volumes mediated the correlations between the abnormal microstructure indicators in WMH region and the decreased memory scores in the SIVD-CI group. Conclusion Patients with SIVD had structural and functional damages in both WMH and NAWM, along with specific gray matter atrophy, which were closely related to memory impairment, especially CA1 atrophy and thalamic atrophy. More importantly, the volumes of some temporomesial regions and the MD values of WMH regions and NAWM may be potentially helpful neuroimaging indicators for distinguishing between SIVD-CI and SIVD-CU patients.
... It is also important to note that there are alternative pathways other than via LC modulation through which the daily practice of heart rate oscillation biofeedback could influence hippocampal volume. For instance, the hippocampus is especially susceptible to impaired cerebrovascular dynamics (Kril et al., 2002;Raz et al., 2007), which may have been affected by the intervention. For instance, if the intervention increased blood flow to the hippocampus, this could potentially influence hippocampal volume (e.g., Fierstra et al., 2010). ...
... In our results (Figures 3aA and 5aA), we observe that pMCI shows an atrophy pattern primarily distributed in CA1 and SUB in early AD, suggesting a similar diffusing pattern from CA1 and SUB to CA2, CA3, and DG (Braak et al., 1993;Braak & Braak, 1991;Fukutani et al., 2000;Gunten et al., 2006;Kril et al., 2002;Lace et al., 2009;Padurariu et al., 2012). Furthermore, our experiment revealed that the medial part of the hippocampal head (SRLM + CA1) exhibited the fastest atrophy in both the left and right hippocampi, with an annual rate of 2.89 ± 5.69% for the left side and 2.96 ± 5.70% for the right side. ...
Article
Full-text available
Increasing evidence has shown a higher sensitivity of Alzheimer's disease (AD) progression by local hippocampal atrophy rather than the whole volume. However, existing morphological methods based on subfield-volume or surface in imaging studies are not capable to describe the comprehensive process of hippocampal atrophy as sensitive as histological findings. To map histological distinctive measurements onto medical magnetic resonance (MR) images, we propose a multiscale skeletal representation (m-s-rep) to quantify focal hippocampal atrophy during AD progression in longitudinal cohorts from the Alzheimer's Disease Neuroimaging Initiative (ADNI). The m-s-rep captures large-to-small-scale hippocampal morphology by spoke interpolation over label projection on skeletal models. To enhance morphological correspondence within subjects, we align the longitudinal m-s-reps by surface-based transformations from baseline to subsequent timepoints. Cross-sectional and longitudinal measurements derived from m-s-rep are statistically analyzed to comprehensively evaluate the bilateral hippocampal atrophy. Our findings reveal that during the early AD progression, atrophy primarily affects the lateral-medial extent of the hippocampus, with a difference of 1.8 mm in lateral-medial width in 2 years preceding conversion (p < .001), and the medial head exhibits a maximum difference of 3.05%/year in local atrophy rate (p = .011) compared to controls. Moreover, progressive mild cognitive impairment (pMCI) exhibits more severe and widespread atrophy in the head and body compared to stable mild cognitive impairment (sMCI), with a maximum difference of 1.21 mm in thickness in the medial head 1 year preceding conversion (p = .012). In summary, our proposed method can quantitatively measure the hippocampal morphological changes on 3T MR images, potentially assisting the pre-diagnosis and prognosis of AD.
... Cerebral small vessel disease has also been reported to increase dementia risk Cannistraro et al., 2019;Kim et al., 2020). Patients with dementia that have microvascular pathology have demonstrated significant loss of hippocampal neurons (Kril et al., 2002), including patients with hereditary cerebral small vessel disease (Yamamoto et al., 2021). Preclinical decrements in vascular function, i.e., changes that do not result in overt disease, may also contribute to dementia etiology. ...
Article
Full-text available
Space exploration requires the characterization and management or mitigation of a variety of human health risks. Exposure to space radiation is one of the main health concerns because it has the potential to increase the risk of cancer, cardiovascular disease, and both acute and late neurodegeneration. Space radiation-induced decrements to the vascular system may impact the risk for cerebrovascular disease and consequent dementia. These risks may be independent or synergistic with direct damage to central nervous system tissues. The purpose of this work is to review epidemiological and experimental data regarding the impact of low-to-moderate dose ionizing radiation on the central nervous system and the cerebrovascular system. A proposed framework outlines how space radiation-induced effects on the vasculature may increase risk for both cerebrovascular dysfunction and neural and cognitive adverse outcomes. The results of this work suggest that there are multiple processes by which ionizing radiation exposure may impact cerebrovascular function including increases in oxidative stress, neuroinflammation, endothelial cell dysfunction, arterial stiffening, atherosclerosis, and cerebral amyloid angiopathy. Cerebrovascular adverse outcomes may also promote neural and cognitive adverse outcomes. However, there are many gaps in both the human and preclinical evidence base regarding the long-term impact of ionizing radiation exposure on brain health due to heterogeneity in both exposures and outcomes. The unique composition of the space radiation environment makes the translation of the evidence base from terrestrial exposures to space exposures difficult. Additional investigation and understanding of the impact of low-to-moderate doses of ionizing radiation including high (H) atomic number (Z) and energy (E) (HZE) ions on the cerebrovascular system is needed. Furthermore, investigation of how decrements in vascular systems may contribute to development of neurodegenerative diseases in independent or synergistic pathways is important for protecting the long-term health of astronauts.
Preprint
Full-text available
Using data from a clinical trial, we tested the hypothesis that daily sessions modulating heart rate oscillations affect older adults’ volume of a region-of-interest (ROI) comprised of adjacent hippocampal subregions with relatively strong locus coeruleus (LC) noradrenergic input. Younger (N=106) and older adults (N=56) completed five weeks of heart rate variability biofeedback. Participants were randomly assigned to one of two daily biofeedback practices: 1) engage in slow-paced breathing (around 10-s per breath or 0.10 Hz) while receiving biofeedback to increase their heart rate oscillations (Osc+); 2) engage in self-selected strategies while receiving biofeedback to decrease their heart rate oscillations (Osc−). Biofeedback did not significantly affect younger adults’ hippocampal volume. Among older adults, the two biofeedback conditions affected volume in the LC-targeted hippocampal ROI differentially with the Osc+ condition showing relatively increased volume and Osc− showing relatively decreased volume. Older adults with greater average spectral power of heart rate at around 0.10 Hz during biofeedback sessions showed greater increases in volume across these regions.
Article
Full-text available
Background: Vascular dementia (VaD) and Alzheimer's disease (AD) are the two most common forms of dementia. Although these two types of dementia have different etiologies, they share some similarities in their pathophysiology, such as neuronal loss and decreased levels of tau protein. We hypothesize that these can have an impact upon the molecular changes in tubulin, precede the neuronal cell loss, and lead to changes in cytoskeletal associated proteins, as documented in both VaD and AD. Objective: We characterized different isotypes of tubulin together with their posttranslational modifications, as well as several microtubule associated proteins (MAPs), such as tau protein, MAP2 and MAP6, all together known as the tubulin code. Methods: We performed western blotting in human brain homogenates of controls and AD and VaD subjects. Results: We report that the levels of different tubulin isotypes differ depending on the dementia type and the brain area being studied: whereas α-tubulin is increased in the temporal lobe of VaD patients, it is decreased in the frontal lobe of AD patients. In VaD patients, the frontal lobe had a decrease in tyrosinated tubulin, which was accompanied by a decrease in tau protein and a tendency for lower levels of MAP2. Conclusion: Our findings highlight distinct changes in the tubulin code in VaD and AD, suggesting a therapeutic opportunity for different dementia subtypes in the future.
Article
Eighty-three brains obtained at autopsy from nondemented and demented individuals were examined for extracellular amyloid deposits and intraneuronal neurofibrillary changes. The distribution pattern and packing density of amyloid deposits turned out to be of limited significance for differentiation of neuropathological stages. Neurofibrillary changes occurred in the form of neuritic plaques, neurofibrillary tangles and neuropil threads. The distribution of neuritic plaques varied widely not only within architectonic units but also from one individual to another. Neurofibrillary tangles and neuropil threads, in contrast, exhibited a characteristic distribution pattern permitting the differentiation of six stages. The first two stages were characterized by an either mild or severe alteration of the transentorhinal layer Pre-alpha (transentorhinal stages I-II). The two forms of limbic stages (stages III-IV) were marked by a conspicuous affection of layer Pre-alpha in both transentorhinal region and proper entorhinal cortex. In addition, there was mild involvement of the first Ammon's horn sector. The hallmark of the two isocortical stages (stages V-VI) was the destruction of virtually all isocortical association areas. The investigation showed that recognition of the six stages required qualitative evaluation of only a few key preparations.
Article
The number of pyramidal neurones in the hippocampal cortex was determined in serially sectioned mesial temporal lobe from brains of 18 mentally normal people; as well as those of 8 demented patients with pathologically confirmed Alzheimer's disease. Normal ageing was accompanied by a gradual loss of neurones, whereas dements' brains showed a much more severe decrease, exceeding that of controls at any age. A high degree of negative exponential correlation was found between the density of neurones/mm3 of cortex and both the number of neurones with neurofibrillary degeneration and the number with granulovacuolar degeneration. The functional significance of the latter changes is thus probably greater than previously assumed, given the diminished population of surviving neurones in which these alterations appear. Both tangles and granulovacuoles demonstrated a stronger propensity for occurring in the posterior half of the hippocampus in demented patients' brains. This would not have been predicted from the relative distribution of neuronal loss in the two halves. The posterior portion of the hippocampus may be considerably more susceptible to the degenerative nerve cell changes prominent in dementia of the Alzheimer type.
Article
Neuronal death following ischemic insults has been thought to reflect necrosis. However, recent evidence from several labs suggests that programmed cell death, leading to apoptosis, might additionally contribute to this death. We have used both in vitro and in vivo models to study the role of apoptosis in ischemic cell death. Some features of apoptosis (TUNEL staining, internucleosomal DNA fragmentation, sensitivity to cycloheximide) were observed following transient focal ischemia in rats. Brief transient focal ischemia was followed by delayed infarction more than 3 days later; this delayed infarction was sensitive to cycloheximide. A cycloheximide-sensitive component of neuronal cell death was also observed in cultured murine neocortical neurons deprived of oxygen-glucose in the presence of glutamate receptor antagonists. This presumed ischemic apoptosis was attenuated by caspase inhibitors, or by homozygous deletion of the bax gene. Neurons may undergo both apoptosis and necrosis after ischemic insults, and thus it may be therapeutically desirable to block both processes.
Article
The examination of neurofibrillary tangles is now recommended for the diagnosis of Alzheimer’s disease as their location and density can distinguish early, intermediate and late disease stages. While the Braak tangle staging protocol can identify these stages, it uses an uncommon silver stain and hippocampal sample. The present study evaluates the Braak protocol using commonly used methods and cases fulfilling either CERAD criteria for Alzheimer’s disease, criteria for dementia with Lewy bodies or without neurological disease. Temporal and occipital cortices from 72 cases were stained using tau immunohistochemistry and the Gallyas and modified Bielschowsky silver stains. The modified Bielschowsky silver stain was equivalent to the Gallyas silver stain for tangle staging. Semiquantitative evaluation of neurofibrillary tangles in the hippocampus and the inferior temporal cortex provided equivalent information to that obtained using the original Braak tangle staging protocol (kappa statistic of 0.97). Comparison of this modification with the CERAD criteria provided moderate agreement (0.51) between diagnostic categories when cases with dementia with Lewy bodies were included, but substantially increased agreement (0.74) when they were excluded. This simplification of the Braak tangle staging protocol is easy to apply, can be readily incorporated into existing CERAD procedures, and helps to distinguish cases with neurofibrillary tangles from those with Lewy bodies.
Article
Eighty-three brains obtained at autopsy from nondemented and demented individuals were examined for extracellular amyloid deposits and intraneuronal neurofibrillary changes. The distribution pattern and packing density of amyloid deposits turned out to be of limited significance for differentiation of neuropathological stages. Neurofibrillary changes occurred in the form of neuritic plaques, neurofibrillary tangles and neuropil threads. The distribution of neuritic plaques varied widely not only within architectonic units but also from one individual to another. Neurofibrillary tangles and neuropil threads, in contrast, exhibited a characteristic distribution pattern permitting the differentiation of six stages. The first two stages were characterized by an either mild or severe alteration of the transentorhinal layer Pre-alpha (transentorhinal stages I-II). The two forms of limbic stages (stages III-IV) were marked by a conspicuous affection of layer Pre-alpha in both transentorhinal region and proper entorhinal cortex. In addition, there was mild involvement of the first Ammon's horn sector. The hallmark of the two isocortical stages (stages V-VI) was the destruction of virtually all isocortical association areas. The investigation showed that recognition of the six stages required qualitative evaluation of only a few key preparations.