+ All Categories
Home > Documents > Cortical Change in Alzheimer’s Disease Detected with a...

Cortical Change in Alzheimer’s Disease Detected with a...

Date post: 21-Jul-2020
Category:
Upload: others
View: 0 times
Download: 0 times
Share this document with a friend
16
We report the first detailed population-based maps of cortical gray matter loss in Alzheimer’s disease (AD), revealing prominent features of early structural change. New computational approaches were used to: (i) distinguish variations in gray matter distribution from variations in gyral patterns; (ii) encode these variations in a brain atlas (n = 46); (iii) create detailed maps localizing gray matter differences across groups. High resolution 3D magnetic resonance imaging (MRI) volumes were acquired from 26 subjects with mild to moderate AD (age 75.8 ± 1.7 years, MMSE score 20.0 ± 0.9) and 20 normal elderly controls (72.4 ± 1.3 years) matched for age, sex, handedness and educational level. Image data were aligned into a standardized coordinate space specifically developed for an elderly population. Eighty-four anatomical models per brain, based on parametric surface meshes, were created for all 46 subjects. Struc- tures modeled included: cortical surfaces, all major superficial and deep cortical sulci, callosal and hippocampal surfaces, 14 ventricular regions and 36 gyral boundaries. An elastic warping approach, driven by anatomical features, was then used to measure gyral pattern variations. Measures of gray matter distribution were made in corresponding regions of cortex across all 46 subjects. Statistical variations in cortical patterning, asymmetry, gray matter distribution and average gray matter loss were then encoded locally across the cortex. Maps of group differences were generated. Average maps revealed complex profiles of gray matter loss in disease. Greatest deficits (20–30% loss, P < 0.001–0.0001) were mapped in the temporo-parietal cortices. The sensorimotor and occipital cortices were comparatively spared (0–5% loss, P > 0.05). Gray matter loss was greater in the left hemisphere, with different patterns in the heteromodal and idiotypic cortex. Gyral pattern variability also differed in cortical regions appearing at different embryonic phases. 3D mapping revealed profiles of structural deficits consistent with the cognitive, metabolic and histological changes in early AD. These deficits can therefore be (i) charted in a living population and (ii) compared across individuals and groups, facilitating longitudinal, genetic and interventional studies of dementia. Introduction Considerable research has focused on uncovering specific patterns of cortical change in Alzheimer’s disease (AD) and other dementias (Friedland and Luxenberg, 1988), schizophrenia (Kikinis et al., 1994; Csernansky et al., 1998), epilepsy (Cook et al., 1994), attention deficit hyperactivity disorder (ADHD) (Giedd et al., 1994), autism (Filipek et al., 1989; Courchesne, 1997) and cortical dysplasias (Sobire et al., 1995) as well as normal development (Thompson et al., 2000b). In AD early neuronal loss occurs in the entorhinal, parahippocampal and temporo-parietal cortex, consistent with the spatial pattern of early perfusion deficits and metabolic change. These deficits mirror the time course of cognitive impairment, proceeding from the entorhinal, temporal and perisylvian association cortices into more anterior regions as the disease progresses. In principle, volumetric magnetic resonance imaging (MRI) scans have sufficient resolution and tissue contrast to track cortical gray matter loss in a living individual. Yet, gyral patterns are extremely variable across subjects, making it difficult to calib- rate individual patterns of gray matter loss against a normative population. It is also hard to determine the average profile of early tissue loss in a group. If 3D profiles of gray matter could be compared, this could be useful (i) for early diagnosis and assessing disease modification in an individual or group and (ii) for understanding how cortical changes relate to the fundamental anatomy of the cortex. This paper addresses these problems. It offers an approach to compare a patient’s cortical anatomy and gray matter distribution with a population-based control image database. The approach is used to resolve group-specific patterns of gray matter distribution and cortical organization. The maps are made by first creating a brain atlas that encodes information on anatomical variability and cortical gray matter distribution in a population (Mazziotta et al., 1995; Thompson et al., 1997, 1998, 2000; Grenander and Miller, 1998). We use a new computational strategy to compute gyral pattern variations across subjects. This approach is central to the study. It was used, rather than a more conventional stereotaxic approach, so that profiles of gray matter loss could be related to the gyral anatomy of each individual and not be confounded by the spatial variability within a stereotaxic template. Measures of gray matter made in each subject could then be averaged across regions of cortex that corresponded anatomically, rather than just stereotaxically. The resulting maps allow measures of gray matter loss to be plotted relative to the gyral anatomy of the cortex. Profiles of gray matter loss (and the spatial variability of the cortical pattern) can then be plotted on average models of the cortex for each group. In these average maps gyral features are well resolved and appear in their mean spatial locations. In this way, profiles of early gray matter loss are mathematically separated from underlying differences in cortical patterns and registration mismatch. Group differences are then displayed visually, using color coded 3D maps. Hypotheses We hypothesized that population-based averaging of anatomy would reveal a region of earliest tissue loss in the left temporal and parietal cortices, with a comparative sparing of the sensori- motor and occipital cortices. We also predicted that the temporo- parietal cortices, specifically the left perisylvian language regions, would exhibit the greatest spatial variability, making it difficult to resolve these early structural changes without specialized approaches to control for such high anatomical variance (Thompson et al., 1997). These new approaches were also designed to allow mapping of cortical asymmetries important in evaluating evidence for the asymmetrical pro- gression of the disease. In this way, true differences in gray matter distribution and cortical metabolism can be distinguished Cortical Change in Alzheimer’s Disease Detected with a Disease-specific Population-based Brain Atlas Paul M. Thompson 1 , Michael S. Mega 1,2 , Roger P. Woods 1 , Chris I. Zoumalan 1 , Chris J. Lindshield 1 , Rebecca E. Blanton 1 , Jacob Moussai 1 , Colin J. Holmes 1 , Jeffrey L. Cummings 2 and Arthur W. Toga 1 1 Laboratory of Neuro Imaging, Department of Neurology, Division of Brain Mapping and 2 Alzheimer’s Disease Center, UCLA School of Medicine, Los Angeles, CA, USA Cerebral Cortex Jan 2001;11:1–16; 1047–3211/01/$4.00 © Oxford University Press 2001. All rights reserved.
Transcript
Page 1: Cortical Change in Alzheimer’s Disease Detected with a ...users.loni.ucla.edu/~thompson/CERCOR2001/cereb_ctx2001.pdf · specially constructed to reflect the average morphology of

We report the first detailed population-based maps of cortical graymatter loss in Alzheimer’s disease (AD), revealing prominent featuresof early structural change. New computational approaches wereused to: (i) distinguish variations in gray matter distribution fromvariations in gyral patterns; (ii) encode these variations in a brainatlas (n = 46); (iii) create detailed maps localizing gray matterdifferences across groups. High resolution 3D magnetic resonanceimaging (MRI) volumes were acquired from 26 subjects with mild tomoderate AD (age 75.8 ± 1.7 years, MMSE score 20.0 ± 0.9) and 20normal elderly controls (72.4 ± 1.3 years) matched for age, sex,handedness and educational level. Image data were aligned into astandardized coordinate space specifically developed for an elderlypopulation. Eighty-four anatomical models per brain, based onparametric surface meshes, were created for all 46 subjects. Struc-tures modeled included: cortical surfaces, all major superficial anddeep cortical sulci, callosal and hippocampal surfaces, 14 ventricularregions and 36 gyral boundaries. An elastic warping approach, drivenby anatomical features, was then used to measure gyral patternvariations. Measures of gray matter distribution were made incorresponding regions of cortex across all 46 subjects. Statisticalvariations in cortical patterning, asymmetry, gray matter distributionand average gray matter loss were then encoded locally across thecortex. Maps of group differences were generated. Average mapsrevealed complex profiles of gray matter loss in disease. Greatestdeficits (20–30% loss, P < 0.001–0.0001) were mapped in thetemporo-parietal cortices. The sensorimotor and occipital corticeswere comparatively spared (0–5% loss, P > 0.05). Gray matter losswas greater in the left hemisphere, with different patterns in theheteromodal and idiotypic cortex. Gyral pattern variability alsodiffered in cortical regions appearing at different embryonic phases.3D mapping revealed profiles of structural deficits consistent withthe cognitive, metabolic and histological changes in early AD. Thesedeficits can therefore be (i) charted in a living population and (ii)compared across individuals and groups, facilitating longitudinal,genetic and interventional studies of dementia.

IntroductionConsiderable research has focused on uncovering specific

patterns of cortical change in Alzheimer’s disease (AD) and other

dementias (Friedland and Luxenberg, 1988), schizophrenia

(Kikinis et al., 1994; Csernansky et al., 1998), epilepsy (Cook

et al., 1994), attention deficit hyperactivity disorder (ADHD)

(Giedd et al., 1994), autism (Filipek et al., 1989; Courchesne,

1997) and cortical dysplasias (Sobire et al., 1995) as well as

normal development (Thompson et al., 2000b). In AD early

neuronal loss occurs in the entorhinal, parahippocampal and

temporo-parietal cortex, consistent with the spatial pattern of

early perfusion deficits and metabolic change. These deficits

mirror the time course of cognitive impairment, proceeding

from the entorhinal, temporal and perisylvian association

cortices into more anterior regions as the disease progresses. In

principle, volumetric magnetic resonance imaging (MRI) scans

have sufficient resolution and tissue contrast to track cortical

gray matter loss in a living individual. Yet, gyral patterns are

extremely variable across subjects, making it difficult to calib-

rate individual patterns of gray matter loss against a normative

population. It is also hard to determine the average profile of

early tissue loss in a group. If 3D profiles of gray matter could be

compared, this could be useful (i) for early diagnosis and

assessing disease modification in an individual or group and

(ii) for understanding how cortical changes relate to the

fundamental anatomy of the cortex. This paper addresses these

problems. It offers an approach to compare a patient’s cortical

anatomy and gray matter distribution with a population-based

control image database. The approach is used to resolve

group-specific patterns of gray matter distribution and cortical

organization. The maps are made by first creating a brain atlas

that encodes information on anatomical variability and cortical

gray matter distribution in a population (Mazziotta et al., 1995;

Thompson et al., 1997, 1998, 2000; Grenander and Miller,

1998).

We use a new computational strategy to compute gyral pattern

variations across subjects. This approach is central to the study.

It was used, rather than a more conventional stereotaxic

approach, so that profiles of gray matter loss could be related to

the gyral anatomy of each individual and not be confounded by

the spatial variability within a stereotaxic template. Measures of

gray matter made in each subject could then be averaged across

regions of cortex that corresponded anatomically, rather than

just stereotaxically. The resulting maps allow measures of gray

matter loss to be plotted relative to the gyral anatomy of the

cortex. Profiles of gray matter loss (and the spatial variability of

the cortical pattern) can then be plotted on average models

of the cortex for each group. In these average maps gyral

features are well resolved and appear in their mean spatial

locations. In this way, profiles of early gray matter loss are

mathematically separated from underlying differences in cortical

patterns and registration mismatch. Group differences are then

displayed visually, using color coded 3D maps.

Hypotheses

We hypothesized that population-based averaging of anatomy

would reveal a region of earliest tissue loss in the left temporal

and parietal cortices, with a comparative sparing of the sensori-

motor and occipital cortices. We also predicted that the temporo-

parietal cortices, specifically the left perisylvian language

regions, would exhibit the greatest spatial variability, making it

difficult to resolve these early structural changes without

specialized approaches to control for such high anatomical

variance (Thompson et al., 1997). These new approaches

were also designed to allow mapping of cortical asymmetries

important in evaluating evidence for the asymmetrical pro-

gression of the disease. In this way, true differences in gray

matter distribution and cortical metabolism can be distinguished

Cortical Change in Alzheimer’s DiseaseDetected with a Disease-specificPopulation-based Brain Atlas

Paul M. Thompson1, Michael S. Mega1,2, Roger P. Woods1,

Chris I. Zoumalan1, Chris J. Lindshield1, Rebecca E. Blanton1,

Jacob Moussai1, Colin J. Holmes1, Jeffrey L. Cummings2 and

Arthur W. Toga1

1Laboratory of Neuro Imaging, Department of Neurology,

Division of Brain Mapping and 2Alzheimer’s Disease Center,

UCLA School of Medicine, Los Angeles, CA, USA

Cerebral Cortex Jan 2001;11:1–16; 1047–3211/01/$4.00© Oxford University Press 2001. All rights reserved.

Page 2: Cortical Change in Alzheimer’s Disease Detected with a ...users.loni.ucla.edu/~thompson/CERCOR2001/cereb_ctx2001.pdf · specially constructed to reflect the average morphology of

from individual or hemispheric differences in cortical organiza-

tion.

Materials and Methods

Subjects and Image Acquisition

Imaging

High resolution 3D MRI volumes were acquired from 26 subjects

diagnosed with mild to moderate AD and 20 elderly control subjects.

Image volumes were acquired on a GE Signa 1.5 T clinical scanner

(Milwaukee, WI). The scans were T1-weighted fast SPGR (spoiled GRASS)

images with a 256 × 256 × 124 imaging matrix. Acquisition parameters

were: TR/TE 14.3/3.2 ms, f lip angle 35°, number of excitations 1, field of

view 25 cm and contiguous 1.2 mm thick slices (no interslice gap)

covering the entire brain. The resulting in-plane pixel resolution was

0.9765 × 0.9765 mm, sufficient for resolving the detailed anatomy of the

cortex.

Diagnosis

The 26 patients met the National Institute of Neurological and Com-

municative Disorders and Stroke/Alzheimer’s Disease and Related

Disorders Association (NINCDS-ARDRA) criteria for AD (McKhann et al.,

1984). In addition, the patients had an acquired persistent decline

involving at least three of the following domains: language, memory,

visuospatial skills, cognition, emotion or personality (Cummings et al.,

1980). Exclusion criteria for all subjects were the presence of a focal

lesion on brain MRI, history of head trauma, past psychiatric history or an

active medical problem.

Demographics

Patients were matched for age (75.8 ± 1.7 years, 14 females/12 males),

educational level (15.2 ± 0.4 years), disease severity and handedness (all

right-handed). Their mean Mini-Mental State Exam score of 20.0 ± 0.9

(maximum score 30) (Folstein et al., 1975) was carefully matched across

the patient cohort to ref lect the mild AD population typically presenting

initially to a clinic (Murphy et al., 1993). The 20 elderly control subjects

were matched with the patients for age, sex, educational level and

handedness (mean age 72.4 ± 1.3 years, 8 females/12 males, mean

educational level 15.4 ± 0.5 years, all right-handed).

Image Alignment and Pre-processing

3D Image Alignment

Imaging data were first aligned to a standard anatomical image template,

specially constructed to ref lect the average morphology of an elderly

population. The construction of this template has been described in

detail (Thompson et al., 2000a). It forms the core of a growing disease-

specific atlas of the brain in AD (Thompson et al., 2000a,b,c; Mega et al.,

1997, 1998, 2000a,b). Brief ly, the average MRI brain template (Fig. 1c)

was constructed to have the average shape and size for a group of elderly

subjects. Specialized approaches for anatomical averaging were used to

generate an average MRI scan with well-resolved cortical features in their

mean spatial locations. The resulting template ref lects the average

morphology of an elderly group of subjects and better ref lects the

anatomy of the subjects in this study than imaging templates based on

young normals (Evans et al., 1994) (Fig. 1a) or post-mortem data

(Talairach and Tournoux, 1988).

Table 1Network of gyral boundaries and cortical surface landmarks

Cortical region Abbreviation

Frontal1 Central sulcus CENT2 Post-central sulcus PoCENT3 Superior frontal sulcus SFS4 Inferior frontal sulcus IFS5 Olfactory sulcus OlfS

Temporal1 Sylvian fissure SYLV2 Superior temporal sulcus STS3 Inferior temporal sulcus ITS4 Collateral sulcus CoS

Parietal1 Intraparietal sulcus IPS

Occipital1 Occipito-temporal sulcus OTS

Marginal1 Inferior frontal margin IFm2 Anterior frontal margin AFm3 Superior frontal margin SFm4 Post–central margin PoCem5 Parietal margin PARm6 Occipital margin OCCm7 Lingual margin LINGm

As major functional interfaces in the brain, these primary sulci and cortical landmarks wereselected because they mark critical gyral and lobar boundaries and extend sufficiently across theexterior brain surface to reflect distributed variations in neuroanatomy across subjects. Allstructures were traced in both brain hemispheres.

Figure 1. Elderly brain template. In this study all 46 scans were aligned to an imaging template based on the anatomy of elderly subjects (c). This template is shown here forcomparison with a widely used average brain image template (ICBM305) based on young normals (a). In the ICBM305 template, which was created by pixel-by-pixel intensityaveraging of 305 young normal subjects’ scans (Evans et al., 1994), notice how anatomical features are not well resolved at the cortex. Cortical variability is represented usingprobability clouds (top left) that describe the frequency of incidence for each gyrus at each stereotaxic voxel, after linear registration and voxel-by-voxel comparison. In a brain template(b) similarly constructed from nine AD patients’ scans the cortical average is also poorly resolved. In contrast, anatomical features are highly resolved, even at the cortex, in theContinuum-Mechanical Brain Template (c), which applies a cortical matching transformation to each brain before intensity averaging (Thompson et al., 2000a). Scans are elasticallyreconfigured into a group mean configuration, using surface-based warping to match 84 surface models (including gyral pattern elements) across all subjects. The intensities of thereconfigured scans are then averaged voxel-by-voxel, after intensity normalization. This produces a group image template with the average geometry and average image intensity forthe group. Elastic transformations are required to resolve cortical features, in their mean configuration, after scans are averaged together (cf. Grenander and Miller, 1998). The data inthis study were aligned to this specially constructed elderly image template, which better reflects elderly anatomy than templates based on young normals.

Figure 2. Computing differences in cortical patterns. Cortical anatomy can be compared, for any pair of subjects (3D models; top left), by computing the 3D deformation field thatreconfigures one subject’s cortex onto another (right panel). In this mapping gyral patterns must also be constrained to match their counterparts in the target brain. To do this, thealgorithm that automatically extracts a 3D model of the cortex provides a continuous inverse mapping from each subject’s cortex to a sphere or plane. A flow field in the flat parameterspace then drives the cortical features into register (c,d). The full mapping (bottom right) can be recovered in 3D space as a displacement vector field that drives cortical points andregions in one brain into precise structural registration with their counterparts in the other brain. Gyral patterns can also be matched across a group of subjects to create averagecortical surfaces. (c) A cortical flat map for a hemisphere of one subject, with the average cortical pattern for the group overlaid (colored lines). (d) The result of warping the individual’ssulcal pattern into the average configuration for the group, using the covariant flow equations (Thompson et al., 2000a). The 3D cortical regions that map to these average locationsare then recovered in each individual subject, as follows. A color code (Thompson and Toga, 1997) representing 3D cortical point locations (e) in this subject is convected along withthe flow that drives the sulcal pattern into the average configuration for the group (d). Once this is done for all subjects, points on each individual’s cortex are recovered (f,g) that havethe same relative location to the primary folding pattern in all subjects. Averaging of these corresponding points results in a crisp average cortex. The transformation fields that mapindividuals onto the group average (h) are stored and used to measure regional variability.

2 Cortical Change in Alzheimer’s Disease • Thompson et al.

Page 3: Cortical Change in Alzheimer’s Disease Detected with a ...users.loni.ucla.edu/~thompson/CERCOR2001/cereb_ctx2001.pdf · specially constructed to reflect the average morphology of

Imaging data were aligned to this template using Automated Image

Registration software (Woods et al., 1993, 1998). An affine (linear)

transformation was applied to individual datasets to transform them into

our common coordinate space. This standard alignment was necessary

for automated extraction of anatomical structure models and generation

of tissue maps.

RF Inhomogeneity Correction

MRI volumes were corrected for potential non-uniformities in MR signal

intensity due to field inhomogeneities in the scanner. An automated

algorithm (Zijdenbos and Dawant, 1994; Sled et al., 1998) derived a 3D

map of these low frequency signal f luctuations and divided by this scalar

field to correct for any errors at the voxel level.

Cerebral Cortex Jan 2001, V 11 N 1 3

Page 4: Cortical Change in Alzheimer’s Disease Detected with a ...users.loni.ucla.edu/~thompson/CERCOR2001/cereb_ctx2001.pdf · specially constructed to reflect the average morphology of

Tissue Classification and Mapping

For all 46 subjects maps of gray matter, white matter and cerebrospinal

f luid (CSF) were generated, so that regional differences in gray matter

could be identified. Brief ly, samples of each tissue class were interactively

tagged to compute the parameters of a Gaussian mixture distribution that

ref lects statistical variability in the intensity of each tissue type (Sled et

al., 1998). A nearest neighbor tissue classifier then assigned each image

voxel to a particular tissue class (gray, white or CSF) or to a background

class (representing extracerebral voxels in the image). The inter/intra-

rater reliability of this protocol and its robustness to changes in image

acquisition parameters have been described previously (Sowell et al.,

1999a,b). Gray matter maps were retained for subsequent analysis.

Cortical Surface Extraction

Next, a high resolution shape representation of the cortex was auto-

matically extracted for each subject, as described previously (Thompson

et al., 1997). This algorithm successively deforms a spherical surface into

the configuration of a given subject’s cortex, resolving the gyral pattern.

The ability of the 3D active surface extraction algorithm (MacDonald et

al., 1993, 1994, 1998) to extract high fidelity surface representations of

the CSF/gray matter and gray/white matter interfaces in high resolution

MR data has been extensively tested and validated in prior studies

(Holmes et al., 1996; Thompson et al., 1997, 1998; MacDonald, 1998).

The resulting cortical model consists of a mesh of discrete triangular

elements that tile the surface. The intensity value at which the gray

matter/CSF interface occurred was determined in 20–30 cortical regions

and the threshold set to the mean. This threshold was independently

validated in each case by comparing the automatically extracted surface

boundaries with the manually determined surface of the gray matter–CSF

boundary, identified at high magnification in the corresponding 3D MR

volumes.

Gyral Pattern Modeling

To determine the patterns of variability for individual regions of cortex,

36 additional cortical structures per brain were traced in all 46 subjects.

These 36 major external fissures and sulci in the brain (Table 1) were

manually outlined on a highly magnified surface-rendered image of each

cortex. Priority was given to biological features whose topological

consistency has been demonstrated across normal populations (Ono et

al., 1990; Le Goualher et al., 1996; MacDonald et al., 1997). Detailed

anatomical criteria were applied as previously set out (Steinmetz et al.,

1989, 1990; Missir et al., 1989; Leonard, 1996; Thompson and Toga,

1997; Thompson et al., 1997; Kennedy et al., 1998) and as in a sulcal atlas

(Ono et al., 1990). In both hemispheres 3D curves were drawn to

represent the superior and inferior frontal, central, post-central,

intraparietal, superior and inferior temporal, collateral, olfactory and

occipito-temporal sulci, as well as the Sylvian fissures. Additional 3D

curves were drawn to represent gyral limits at the interhemispheric

margin1 (Thompson et al., 1997). Stereotaxic locations of contour points

derived from the data volume were re-digitized to produce 36 uniformly

parameterized cortical contours per brain, representing the primary gyral

pattern of each subject (Thompson et al., 1997; Thompson and Toga,

1998).

3D Cortical Surface Averaging

Transforming individual data into a standardized coordinate space

removes differences in overall brain size. Nonetheless, substantial

anatomical variability remains, especially at the cortex, due to individual

differences in gyral patterning (Steinmetz et al., 1989, 1990; Thompson et

al., 1996b) and disease-related atrophy (Mega et al., 1998). Information

was stored on individual cortical differences by (i) creating average

geometric models for the cortex and then (ii) measuring individual

deviations from the group average using 3D displacement maps (Fig. 2)

(Thompson et al., 1996a,b, 1997; Thompson and Toga, 1997; Davatzikos

et al., 1996; Mega et al., 1998; Csernansky et al., 1998). These

displacement maps relate points on an individual subject’s cortex to

corresponding points on an average cortical model. Average cortical

model construction has been described previously (Ge et al., 1995;

Collins et al., 1996; Drury and Van Essen, 1997; Thompson et al., 1997,

2000a; Fischl et al., 1999) (see Fig. 2 and Appendix for a summary).

To allow point-to-point cortical averaging, each subject’s cortical

model is converted to a ‘f lat map’ as previously described (Thompson et

al., 2000a). To ensure that each subject’s f lat map can also be converted

back into a 3D cortical model, cortical surface point position vectors in

3D stereotaxic space were represented on the f lat map using a color

code as described previously (Thompson et al., 2000a). This forms an

image of the parameter space in RGB color image format (Fig. 2e). By

carrying a color code (that indexes 3D locations) along with the vector

f low that aligns each individual with the average folding pattern (Fig.

2c,d), information can be recovered at a particular location in the average

folding pattern (Fig. 2f) specifying the 3D cortical points mapping each

subject to the average. The resulting mapping is guaranteed to average

together all points falling on the same cortical locations across the set of

brains and ensures that corresponding features are averaged together

(Fig. 3). It can also be determined which regions of the cortex show

the greatest variability in structure. By using the color code (Fig. 2f) to

identify original cortical locations in 3D space (Fig. 2g), displacement

fields were recovered mapping each subject into gyrus-by-gyrus cor-

respondence with the average cortex (Fig. 4). Anatomical variability was

defined at each point on the average cortical surface as the root mean

square (r.m.s.) magnitude of the 3D displacement vectors assigned to

each point in the surface maps from individual to average (Thompson et

al., 1996a,b, 1997). This measure captures how individuals deviate from

the group average anatomy after taking gyral pattern variations into

account. The resulting variability pattern was visualized as a color coded

map (Fig. 5).

Gray Matter Averaging and Statistical Comparisons of Gray

Matter Distribution

Gray Matter Quantification

Given that the deformation maps associate cortical locations with the

same relation to the primary folding pattern across subjects, a local

measurement of gray matter density was made in each subject and

averaged across equivalent cortical locations. To quantify local gray

matter, we used a measure termed ‘gray matter density’, which has been

used in prior studies to compare the spatial distribution of gray matter

Figure 3. Average cortex in AD. The average cortical surface for the group is shown (bottom row) as a graphically rendered surface model. If sulcal position vectors are averagedwithout aligning the intervening gyral patterns (top), sulcal features are not reinforced across subjects and a smooth average cortex is produced. By matching gyral patterns acrosssubjects before averaging, a crisper average cortex is produced (bottom row). Sulcal features that consistently occur across all subjects appear in their average geometricconfiguration.

Figure 4. Matching an individual’s cortex to the average cortex. 3D variability patterns across the cortex are measured by driving individual cortical patterns into local correspondencewith the average cortical model. (a) How the anatomy of one subject (brown surface mesh) deviates from an average cortical model (white) after affine alignment of the individualdata. (b) The deformation vector field required to reconfigure the gyral pattern of the subject into the exact configuration of the average cortex. The transformation is shown as a flowfield that takes the individual’s anatomy onto the right hemisphere of the average cortex (shown as a blue surface mesh). The largest amount of deformation is required in the temporaland parietal cortex (pink, large deformation). Details of the 3D vector deformation field (b, inset) show the local complexity of the mapping. Storage of these mappings allowsquantification of local anatomical variability.

Figure 5. Statistical map of average gray matter loss in AD (n = 46). Based on averaging and comparing gray matter measurements across equivalent regions of cortex in all 46subjects, this statistical field reflects whether the average gray matter is reduced in patients (average of 26 subjects) relative to controls (average of 20 subjects). The significance ofthis reduction at each cortical location is shown. Severe, more localized reductions are visualized in the temporal lobe and temporo-parietal cortex. This profile of gray matter lossmirrors the anatomical distribution of early perfusion deficits and metabolic change in mild to moderate AD.

4 Cortical Change in Alzheimer’s Disease • Thompson et al.

Page 5: Cortical Change in Alzheimer’s Disease Detected with a ...users.loni.ucla.edu/~thompson/CERCOR2001/cereb_ctx2001.pdf · specially constructed to reflect the average morphology of

Cerebral Cortex Jan 2001, V 11 N 1 5

Page 6: Cortical Change in Alzheimer’s Disease Detected with a ...users.loni.ucla.edu/~thompson/CERCOR2001/cereb_ctx2001.pdf · specially constructed to reflect the average morphology of

across subjects (Sowell et al., 1999a,b; Thompson et al., 2000d). This

measures the proportion of gray matter in a small region of fixed radius

around a point. In these prior studies, however, it was assumed that gray

matter occurring at the same stereotaxic location came from equivalent

anatomical regions across subjects. Given the large anatomical variability

in some cortical regions (Fig. 5), especially in a diseased population, this

assumption is often violated. To avoid this potential methodological error

we employed elastic maps to associate equivalent gyri across both

populations. We were thus able to average gray matter density across

corresponding cortical regions and plot the results on the continuum-

mechanical average AD cortex (Fig. 3). Brief ly, at each cortical point a

sphere of radius 5 mm was made, centered at that point. By reference to

the gray matter maps derived from the tissue classification approach

described above, the proportion of gray matter pixels relative to the total

number of pixels in this sphere was computed and stored as a map of gray

matter densities across the cortex. Because this measure is just another

cortical attribute that can be aligned across subjects, and with the mean

gyral pattern for the group (Figs 3,4), the average gray matter density was

computed across subjects for each cortical point in each group average.

Maps of average gray matter loss were also created by comparing the

average maps from the diseased and control groups. Finally, information

was stored on the variability in gray matter density at equivalent cortical

locations within and across groups. This statistical data within each group

allowed the observed profiles of average gray matter difference to be

calibrated against a variance measure for the index. This variance measure

allowed the significance of local gray matter reductions to be assessed. A

field of test statistics was attached to the average surface for the diseased

group to determine the local statistical significance of the hypothesized

gray matter loss. Finally, a localized test for gray matter loss in the

temporo-parietal cortex was applied, a region where greatest neuronal

loss was hypothesized at this early stage of the disease.

Computer Platform

All algorithms were written in C and executed on Silicon Graphics O2

R10000 workstations running IRIX 6.5, except for the algorithms for

cortical extraction and matching, which were parallelized and executed

on a networked cluster of 14 workstations and a Silicon Graphics

RealityMonster with 32 internal processors.

Results

Cortical Gray Matter Distribution and Disease–Related

Gray Matter Loss

Figure 5 shows a surface-based probability field that indicates the

regional significance of gray matter loss across the cortex in the

entire AD cohort. Red (P < 0.005) denotes brain regions where

the average gray matter index is significantly less2 in the AD

cohort than in the control group. All averages and comparisons

are made across corresponding areas of cortex, defined by gyral

pattern matching (Fig. 4). Given these statistics, two types of

inference are possible. First, the a priori hypothesis of gray

matter loss in the temporal and parietal cortex was confirmed.

There was also evidence for a region of maximal loss throughout

the lateral temporal surface and the parietal operculum

bilaterally (P < 0.001–0.00012).

A pervasive left greater than right hemisphere reduction in

gray matter was found (with up to 20–30% loss locally; see Fig.

7), consistent with the suggestion from metabolic studies

(Loewenstein et al., 1989) that the left hemisphere is, on

average, more severely affected at this stage of the disease. The

occipital cortices were comparatively spared bilaterally, as were

the sensorimotor cortices (0–5% loss, P > 0.05). There was also

severe gray matter loss (20–30%, P < 0.001–0.0001) in the

middle frontal gyrus, in the vicinity of areas 9 and 46 (Rajkowska

and Goldman-Rakic, 1995). We further investigated whether

the regions of more significant gray matter loss ref lected a

correspondingly greater average reduction in the local gray

matter index (Fig. 7). This was important, as a greater sig-

nificance value can result either from (i) a genuinely greater

percent reduction in the mean gray matter in AD or (ii) a local

reduction in the variance of the gray matter index across the

group, which translates into a greater detection sensitivity.

Interestingly, a map of the percentage reduction in average gray

matter (Fig. 7) followed approximately the same anatomical

pattern, suggesting that there is indeed a hierarchy in the

severity of gray matter loss at this stage of the disease, rather than

a f luctuation in the local power of the statistical model to detect

it. Again, the temporal and temporo-parietal cortex exhibited

severe (10–30%) reductions in gray matter. This contrasted with

a comparative sparing of the superior margins of the central and

post-central gyri and occipital poles (0–5% loss). Although

diffuse gray matter loss is likely to occur across the majority of

the cortex, it is interesting that the superior central and

post-central gyri and occipital poles show very little reduction in

gray matter when adjacent posterior temporal cortex and the

parietal operculum are severely affected, in both the percentage

loss and statistical anatomical maps.

Cortical Pattern Variability

As a by-product of the gray matter analysis, maps revealing the

magnitude and directional biases of 3D normal cortical

variability are shown in Figures 8 and 9. For each cortical region

principal directions emerged in which the magnitude of normal

cortical variability was greatest (Fig. 9). The overall magnitude of

variability was also highly heterogeneous. In the control subjects

(n = 20; Fig. 8) variability values rose from 4–5 mm in the

primary motor cortex to localized peaks of maximum variability

in the posterior perisylvian zones and superior frontal

association cortex (12–14 mm). The primary sensory and motor

areas showed a localized invariance relative to all other regions of

the cortex, with bilateral r.m.s. variability values of 2–5 mm at

the central sulcus rising only to 6–9 mm at the post-central

sulcus in both brain hemispheres.

System-specific Variability Patterns

Extremely low variability values in the motor cortex (2–5 mm)

rose with the transition anteriorly from motor area 4 to

Figure 6. 3D cortical variability (n = 26, AD patients). The profile of variability across the cortex is shown, after differences in brain orientation and size are removed. The followingviews are shown: oblique frontal, frontal, right, left, top, bottom. Extreme variability in the posterior perisylvian zones and superior frontal association cortex (12–14 mm, red) contrastswith the comparative invariance of the primary sensory, motor and orbitofrontal cortex (2–5 mm, blue). Models are orthographically projected onto a coordinate grid to facilitatecomparisons with data from functional and metabolic studies (Mega et al., 2000).

Figure 7. Map of average gray matter loss in AD expressed as a percentage of average control values (n = 46). This map expresses the same data as Figure 5 as a percentagereduction in the measurement of gray matter when equivalent cortical regions are averaged and compared between AD patients and controls. As hypothesized, pervasive left greaterthan right reductions are mapped. The percentage reduction in average gray matter followed approximately the same anatomical pattern as the significance map, suggesting thatthere is indeed a hierarchy in the severity of gray matter loss at this stage of the disease, rather than a fluctuation in the local power of the statistical model to detect it. Again, thetemporal and temporo-parietal cortex exhibited severe (10–30%) reductions in gray matter. This contrasted with a comparative sparing of the superior margins of the central andpost-central gyri and the occipital poles (0–5% loss).

6 Cortical Change in Alzheimer’s Disease • Thompson et al.

Page 7: Cortical Change in Alzheimer’s Disease Detected with a ...users.loni.ucla.edu/~thompson/CERCOR2001/cereb_ctx2001.pdf · specially constructed to reflect the average morphology of

Cerebral Cortex Jan 2001, V 11 N 1 7

Page 8: Cortical Change in Alzheimer’s Disease Detected with a ...users.loni.ucla.edu/~thompson/CERCOR2001/cereb_ctx2001.pdf · specially constructed to reflect the average morphology of

8 Cortical Change in Alzheimer’s Disease • Thompson et al.

Page 9: Cortical Change in Alzheimer’s Disease Detected with a ...users.loni.ucla.edu/~thompson/CERCOR2001/cereb_ctx2001.pdf · specially constructed to reflect the average morphology of

Cerebral Cortex Jan 2001, V 11 N 1 9

Page 10: Cortical Change in Alzheimer’s Disease Detected with a ...users.loni.ucla.edu/~thompson/CERCOR2001/cereb_ctx2001.pdf · specially constructed to reflect the average morphology of

pre-frontal association cortex (see Fig. 8a). Peak variability

values (12–14 mm) occurred in the anterior frontal association

cortex on the left and throughout the middle frontal gyrus on

the right, where Brodmann area 46 is consistently located

(Brodmann, 1909; Rajkowska and Goldman-Rakic, 1995). In

these regions of frontal cortex hemispheric differences in gyral

organization are typical (Malobabic et al., 1993). Moving

inferiorly, intermediate variability values (6–10 mm) over the

inferior prefrontal convexity fell with the transition to the

orbitofrontal cortex, where the gyral pattern is highly conserved

across subjects (2–5 mm variability). More laterally, the posterior

frontal cortex, including territory occupied by Broca’s area, also

displayed intermediate variability (6–10 mm). Temporal lobe

variability rose from 2–3 mm in the depths of the Sylvian

cisternae to 18 mm at the posterior limit of the inferior temporal

sulcus in both brain hemispheres (Fig. 8a). This suggests that the

region of maximal variability in the human temporal cortex may

lie posterior to the region of highest variability observed by

Novikov and Podcherednik in primary auditory cortex (Novikov

and Podcherednik, 1992). Furthermore, in the vicinity of the

angular gyrus the 3D r.m.s. variability of the inferior temporal

sulci was substantially greater on the left (12–14 mm) than the

right (10–12 mm). This left greater than right variability pat-

tern was also displayed by the superior temporal sulcus, the

supramarginal gyrus and the posterior ascending ramus of the

Sylvian fissure, which was also considerably more variable on

the left (12–14 mm), where Wernicke’s area is situated, than on

the right (6–10 mm). These findings of asymmetrical variability

support earlier hypotheses by Steinmetz and co-workers, who

examined 2-dimensional sagittal projections of the Sylvian

fissure (Steinmetz et al., 1990).

Tensor Maps Reveal Directional Biases in Cortical Variability

For each region of cortex clear directional biases emerged in the

principal directions of gyral pattern variability (Fig. 9a,b). Gyral

patterns did not vary equally in all directions and the statistical

distribution that describes the location of a cortical region in

space was elongated in a particular direction, which also varied

locally across the cortex. To visualize this, cortical variations

were modeled as vector field displacements of an average

cortical model and ellipsoids of constant probability density

were computed for positions of cortical regions (relative to the

average cortex). Figure 9c shows the shape of a 3D Gaussian

distribution fitted at each point on the average normal cortex,

ref lecting the cross-subject variation of points from equivalent

gyral regions. The shape of this distribution at each cortical

point is described by the covariance tensor of the 3D distribu-

tion. Its value determines a set of nested ellipsoids that represent

confidence limits for the locations of corresponding anatomical

points in stereotaxic space (Thompson et al., 1997; Cao and

Worsley, 2000). These ellipsoids (Fig. 9c) are colored by the

determinant of the covariance tensor, for which larger values

(pink) represent greater 3D variability and small values (blue)

represent regions whose morphology is highly conserved across

subjects.

Anatomical variations in the temporo-parietal regions

displayed the greatest anisotropy, with a strong tendency to vary

in a plane oriented upwards at a 45° angle to the horizontal plane

(see Fig. 9). In several cortical regions the principal directions of

variability (along which the glyphs are elongated in Fig. 9) were

approximately orthogonal to the primary gyral pattern. This

directional trend was similar in some respects to the torquing, or

petalia, which causes cortical regions in the right hemisphere to

be situated slightly anterior to their counterparts on the

left (Galaburda and Geschwind, 1981; Bilder et al., 1994). The

region of highly anisotropic variability was strongly localized to

the temporo-parietal cortex and did not extend anteriorly into

the post-central and central gyri. A marked anatomical division

occurred at the post-central gyrus, where variability was

reduced and was spatially more isotropic. The component of

variability normal to the average cortex was greatest at the

temporal poles, where gyral patterns are relatively stable and

variations in temporal lobe size may dominate. Importantly, this

directional cortical variability is controlled by surface matching

within the continuum-mechanical atlas, thereby allowing accur-

ate maps of disease-related gray matter loss to be constructed

(Figs 5,7).

Cortical Pattern Asymmetry

Figure 10 illustrates the group average patterns of cortical

asymmetry, highlighting regional trends. In a previous study we

found Sylvian fissure asymmetry to be significantly greater in

AD (P < 0.05) than in controls matched for age, gender and

handedness (Thompson et al., 1998). Although these asym-

Figure 8. 3D cortical variability (n = 20, normal elderly subjects). The profile of variability across the cortex is shown after differences in brain orientation and size were removed. Thefollowing views are shown: oblique frontal, frontal, right, left, top, bottom. Again, extreme variability in the posterior perisylvian zones and superior frontal association cortex (12–14mm, red) contrasts with the comparative invariance of the primary sensory, motor and orbitofrontal cortex (2–5 mm, blue). The region of maximal variability, in the temporal cortex, istightly linked with the location of human visual area MT (or V5) (Watson et al., 1993). Extreme caution is therefore necessary when referring to activation foci here using stereotaxiccoordinates. The overall profiles of variation also corroborate recent volumetric findings based on a fine scale parcellation of the cortex (Kennedy et al., 1998), with greatermorphological individuality in phylogenetically more recent cortical regions.

Figure 9. Tensor maps reveal directional biases in normal cortical variability (n = 20). Tensor maps can be used to visualize these complex patterns of gyral pattern variation at thecortex. The maps are based on the group of 20 elderly normal subjects. Color distinguishes regions of high variability (pink) from areas of low variability (blue). In (a) and (b) ellipsoidalglyphs indicate the principal directions of variation: they are most elongated along directions where there is greatest anatomical variation across subjects. Each glyph represents thecovariance tensor of the vector fields that map individual subjects onto their group average anatomical representation. The resulting information can be leveraged to distinguish normalfrom abnormal anatomical variants using random field algorithms and can define statistical distributions for feature labeling at the cortex (Le Goualher et al., 1999; Vaillant andDavatzikos, 1999). (c) Probabilistic confidence limits on normal anatomical variation: tensor field representation. Again, tensor maps reveal the preferred directions of cortical variation,after sulcal pattern correspondences are taken into account. Variability is greatest in the temporo-parietal cortex. Since cortical variations are modeled as vector field displacementsof an average cortical model, ellipsoids of constant probability density can be computed across cortical regions (relative to an average cortex). These probability fields are obtained bysingular value decomposition, or Cholesky factorization, of the local covariance tensor (Thompson et al., 1996a; Cao and Worsley, 2000). Confidence ellipsoids are shown, colored bythe determinant of the covariance tensor, which measures the magnitude of anatomical variability at each location.

Figure 10. Population-based maps of cortical pattern asymmetry. Averaging of cortical patterns across subjects (n = 20, controls) reveals fundamental features in the profile ofasymmetry across the normal human cortex. The marked brain asymmetry in the temporo-parietal cortex is clearly apparent, mapping its average magnitude in a population. Basedon the average models for each cortical sulcus, asymmetry can be quantified locally, in 3D, revealing patterns not apparent in the cortical anatomy of an individual. Asymmetry iscalculated based on 3D displacement maps, which subtract gyral models from mirror images of their counterparts in the opposite hemisphere.

10 Cortical Change in Alzheimer’s Disease • Thompson et al.

Page 11: Cortical Change in Alzheimer’s Disease Detected with a ...users.loni.ucla.edu/~thompson/CERCOR2001/cereb_ctx2001.pdf · specially constructed to reflect the average morphology of

metries are not apparent in every individual, a localized region

can be clearly defined in which major asymmetrical trends are

present (Fig. 10). Severe asymmetry exhibited by the posterior

Sylvian fissure (up to 10 mm) contrasted with negligible

asymmetry in the frontal, parietal and occipital cortex (1–2 mm).

The group average anatomy (Fig. 10) shows the average Sylvian

fissure terminating more posteriorly (P < 0.0002) and oriented

more horizontally on the left than the right, corroborating post-

mortem measurements of the planum temporale (Geschwind

and Levitsky, 1968, Witelson and Kigar, 1992; Galaburda, 1995).

The average right Sylvian fissure also shows an upward turn at its

posterior limit (Fig. 10) and is anterior to the posterior limit on

the left (Thompson et al., 1998).

Surprisingly, these average asymmetries continued anteriorly

into the primary somatosensory cortex and posteriorly into the

inferior temporal cortex (Fig. 10). In Figure 10 the right terminal

rami of the superior and inferior temporal sulci, as well as the

posterior ascending ramus of the Sylvian fissure, were up to

15 mm anterior to their counterparts on the left. This asymmetry

continues into the post-central cortex, with the posterior bank of

the post-central gyrus thrust forward by 8–9 mm on the right

compared to the left (Fig. 10). This asymmetry seems not to be

explainable by the asymmetrical size of the left parietal

operculum, which is larger in most cases on the left (Steinmetz

et al., 1990). The asymmetrical region also covers the territory

occupied by the supramarginal gyrus, which surrounds the

terminal ascending ramus of the Sylvian fissure in both brain

hemispheres. The profile of asymmetry extends caudally across

the planum parietale (Jäncke et al., 1994) and across the lateral

convexity of the cortex into the superior and inferior temporal

gyri, where 3D variation reaches a peak of 14 mm and where

several stereotyped variations in structure have been identified

(Steinmetz et al., 1990; Leonard, 1996).

DiscussionBy averaging cortical features in an AD population and matched

elderly controls, striking profiles of gray matter loss, anatomical

variation and cerebral asymmetry can be identified. Severe

reductions in gray matter (up to 30% loss) were observed across

the lateral temporal surfaces in the AD cohort. These deficits

were also clearly found in the temporo-parietal cortices

bilaterally. Patterns of left greater than right gray matter loss also

became apparent, with severe gray matter loss observed

bilaterally in the vicinity of Brodmann areas 9 and 46, regions

of increased synaptic loss and β-amyloid protein deposition

(Clinton et al., 1994). There was also a comparative sparing of

the superior post-central and central gyri and the occipital poles

(0–5% loss, P < 0.05). This pattern is consistent with preser-

vation of sensorimotor and visual function at this stage of the

disease, at the same time as perfusion and metabolic deficits

pervade in higher order association cortices.

Hemispheric Differences

Interestingly, patterns of greater gray matter loss in the left

hemisphere corroborate earlier reports (Loewenstein et al.,

1989) of predominant left hemisphere metabolic dysfunction in

mild to moderate AD, when cerebral glucose utilization is

measured by positron emission tomography (PET). Structural,

perfusion and metabolic studies suggest that the left hemisphere

may be more susceptible to neuronal loss, instead of the

alternative explanation that equivalent neuronal loss may result

in greater functional deficits on one side, due to asymmetrical

cortical organization. Greatest gray matter loss in the temporo-

parietal cortex may underlie the prominent temporal-parietal

hypometabolism that is consistently found at this stage of AD,

often asymmetrically (Friedland and Luxenberg, 1988; Johnson

et al., 1998). Although the focus of this study was to determine

patterns of gray matter loss in vivo, immunocytochemical

studies have reported between 11 and 50% synaptic loss in

the superior temporal and inferior parietal cortices, with a

comparative sparing of occipital cortices (cf. Figs 5,7). Relatively

greater atrophy is often reported in the temporal lobe relative

to overall cerebral volume (Murphy et al., 1993). The early

progression of AD pathology into the parietal and frontal

association cortices suggests a degeneration of synaptically

linked cortical pathways, and this pattern correlates with symp-

toms of memory impairment, aphasias, apraxias, personality

changes and spatial deficits (Roberts et al., 1993). Interestingly,

gray matter loss at autopsy is predominantly cortical in

Alzheimer’s patients under 80 years of age (Hubbard and

Anderson, 1981), when volumes of subcortical nuclei are not

significantly different between patients and controls (De La

Monte, 1989). Nonetheless, atrophy of the amygdala and basal

nuclei (Cuénod et al., 1993) may ultimately be followed by

alterations in thalamic nuclei (Jernigan et al., 1991), induced

perhaps by degeneration of their cortical projection areas.

Profiles of Tissue Loss

While the lateral temporal and parietal cortices exhibit diffuse

gray matter loss, some regions of the central and paracentral

cortex appear to have several foci of average gray matter loss in

territory that is otherwise comparatively spared (Fig. 7). Gray

matter loss within a gyrus may be a multifocal process (as, for

example, the discrete lesions in vascular dementia) or may occur

rather uniformly within individual gyri. Clearly, some features

occur at small spatial scales in both the statistical (P value) maps

and the average loss maps. This multifocal effect does not appear

to be attributable to sampling error in estimating the variance for

the gray matter measure, as these variance values are spatially

quite homogeneous. Structural and functional features with a

spatial scale smaller than a gyrus may begin to be resolved if data

from corresponding gyri are better aligned across subjects when

averaging features from a population (Thompson et al., 2000a;

Zeineh et al., 2000). Conversely, gyral features may be blurred

out (cf. Fig. 3) when these correspondences are not taken into

account (Evans et al., 1994). We did not hypothesize this multi-

focal effect in advance, so we did not test for its significance

specifically. Longitudinal studies may allow us to better

understand the scale and consistency of these localized changes

over time and may reveal whether gray matter loss is an inher-

ently diffuse or multifocal process within individual cortical

gyri.

Advantages of Gray Matter Maps

Cortical gray matter is lost in AD in a pattern that is temporally

stereotyped and, initially, regionally specific. By resolving this

pattern across the cortex, a detailed evaluation of degenerative

change can be made in living populations. Conventional

volumetric analysis of MRI data shows substantial overlap in both

lobar volumes and gray matter measures between patients and

controls, often because of difficulties in identifying equivalent

areas of cortex. Overall structure volumes also display

considerable variability. High dimensional registration (i.e.

elastic matching) of cortical maps offers a solution to this

difficulty, in that local measurements of gray matter can be

calibrated against a local measure of tissue variance. Large

Cerebral Cortex Jan 2001, V 11 N 1 11

Page 12: Cortical Change in Alzheimer’s Disease Detected with a ...users.loni.ucla.edu/~thompson/CERCOR2001/cereb_ctx2001.pdf · specially constructed to reflect the average morphology of

differences in cortical organization are also readily accom-

modated.

Cortical Pattern Matching

The goal of the cortical matching procedure is to bring cortical

regions into correspondence, so that data from corresponding

regions can be averaged together across subjects. Without a

procedure to align cortical structures, such as the one described

in this paper, an averaging procedure applied voxel-by-voxel in

stereotaxic space does not always average data from the same

region of cortex and, in principle, data from the temporal cortex

of some subjects could be averaged with data from the frontal

cortex of other subjects. The gyral matching procedure alleviates

this problem to a degree, although it does not solve it completely.

Gyral matching does not guarantee that data from corresponding

cytoarchitectonic regions will be averaged together. However,

many functional regions of the cortex defined by PET and

functional MRI (Watson et al., 1993), as well as many cyto-

architectonic regions (Rademacher et al., 1993), bear a

consistent relationship to macroanatomical landmarks of the

gyral pattern. The degree to which cortical pattern matching

reduces architectonic and functional variation can be evaluated

by quantifying residual variability of functional or cellular

landmarks after normalizing gross anatomical features

(Rajkowska and Goldman-Rakic, 1995; Van Essen and Drury,

1997; Fox et al., 1999; Geyer et al., 2000). Differences in the

topological layout of architectonic regions within the cortical

sheet ultimately preclude the mapping of discrete cortical

regions from one subject to another, so an important inter-

mediate goal has been to identify and match a comprehensive

network of sulcal and gyral elements which are consistent in

their incidence and topology across subjects (Ono et al., 1990;

Rademacher et al., 1993; Thompson et al., 1996a, 1997). While

gyral matching substantially reduces the variability in cortical

organization across subjects, in the future functional and

architectonic landmarks may be definable in vivo that better

guarantee matching of the cortical mantle from one subject to

another in population studies (Dumoulin et al., 2000).

At this stage, the pathological burden of AD may be greater in

terms of functional deficits, and synaptic loss, in the hetero-

modal cortex than in the idiotypic cortex. In our prior studies

AD patients exhibited significantly greater asymmetry and

structural variability in the deep perisylvian cortex, relative to

controls matched for age, gender, educational level and

handedness (P < 0.05) (Thompson et al., 1998). Clear differences

in both AD cortical variation and gray matter distribution suggest

the need for disease-specific brain atlases that better ref lect the

disease-related anatomy of patients and calibrate individual loss

against statistical data from normative populations.

Emerging Patterns

In both groups anatomical features emerged that are not

observed in individual representations due to their considerable

variability. As shown in Figure 10, the marked anatomical

asymmetry in the posterior perisylvian cortex (Geschwind and

Levitsky, 1968) extends rostrally into the post-central cortex.

The posterior bank of the post-central gyrus is thrust forward by

8–9 mm on the right compared with the left (Fig. 10). This

asymmetry extends caudally across the lateral convexity into the

superior and inferior temporal cortex. As shown by averaging

models of ventricular anatomy (Thompson et al., 2000d), this

asymmetrical trend penetrates subcortically into the occipital

horns of the lateral ventricles, but not into adjacent parieto-

occipital and calcarine cortex (Thompson et al., 1998). In

contrast with existing brain atlases based on a single brain

hemisphere (Talairach and Tournoux, 1988), population-based

atlases encode information on asymmetry and its group

variation, so that departures from normal patterns in individuals

or groups can be identified (Thompson et al., 1997; Thirion et

al., 1998; Thompson and Toga, 1998; Cao and Worsley, 2000).

There is a substantial literature on Sylvian fissure cortical surface

asymmetries (Eberstaller, 1884; Cunningham, 1892; Geschwind

and Levitsky, 1968; Davidson and Hugdahl, 1994) and their

relation to functional lateralization (Strauss et al., 1983),

handedness (Witelson and Kigar, 1992), language function

(Davidson and Hugdahl, 1994), asymmetries of associated

cytoarchitectonic fields (Galaburda and Geschwind, 1981) and

their thalamic projection areas (Eidelberg and Galaburda, 1982),

However, no prior reports have mapped the asymmetry profile

across the cortex in three dimensions. These localized patterns

of asymmetry in cortical morphology clearly have multiple

determinants. We previously found Sylvian fissure asymmetry to

be significantly greater in AD patients than in controls matched

for age, gender, educational level and handedness (P < 0.05)

(Thompson et al., 1998), suggesting that AD pathology asym-

metrically disrupts the anatomy of the temporo-parietal cortex.

The improved ability to localize asymmetries of cortical

organization or tissue loss in a group atlas presents opportunities

to analyze diseases with asymmetrical progression, including

different stages of AD, and to map hypothesized alterations in

cortical and hippocampal asymmetry in disease states such as

schizophrenia (Falkai et al., 1992; Kikinis et al., 1994; Kulynych

et al., 1996; Csernansky et al., 1998).

Population-based Brain Templates

From a practical standpoint, approaches for anatomical averag-

ing also provide an average anatomical image template to repres-

ent a particular clinical group. In contrast to earlier studies, we

matched cortical patterns across subjects to resolve fundamental

anatomical features across a group. Similar approaches are

under active development to create average brain repres-

entations for the macaque (Grenander and Miller, 1998) and for

individual structures such as the corpus callosum (Gee et al.,

1995; Davatzikos, 1996), central sulcus (Manceaux-Demiau et

al., 1998), cingulate and paracingulate sulci (Paus et al., 1996),

hippocampus (Haller et al., 1997; Csernansky et al., 1998; Joshi

et al., 1998) and for transformed representations of the human

and macaque cortex (Drury and Van Essen, 1997; Grenander and

Miller, 1998; Fischl et al., 1999). The resulting averages provide

templates in which multimodality brain maps can be integrated

(Mazziotta et al., 1995; Toga and Thompson, 1998). The prob-

abil-istic information they contain can also guide Bayesian

approaches for automatically identifying anatomical structures

(Gee et al., 1995; Mangin et al., 1995; Royackkers et al., 1996;

Pitiot et al., 2000). Finally, these probabilistic atlases can

constrain the search space for activations in functional imaging

experiments (Dinov et al., 2000).

A group-specific atlas of the brain in early AD enables

functional, metabolic and tissue distribution data to be analyzed

in an anatomical framework that ref lects AD morphology. The

effects of morphological variation can also be controlled.

However, the strategy described here is applicable, in principle,

to any population. Since AD is a progressive disease, a

homogeneous patient group was selected for this study, matched

for age and educational level, at a stage in the disease when

patients often present for initial evaluation and where MR, PET

12 Cortical Change in Alzheimer’s Disease • Thompson et al.

Page 13: Cortical Change in Alzheimer’s Disease Detected with a ...users.loni.ucla.edu/~thompson/CERCOR2001/cereb_ctx2001.pdf · specially constructed to reflect the average morphology of

and SPECT scans may have maximal diagnostic value. By

expanding the underlying patient database and stratifying the

population according to different criteria, atlases to represent

the more advanced stages of AD, or other clinically defined

groups, could also be developed.

Longitudinal Studies

Longitudinal studies, in which a cohort of subjects is scanned

repeatedly over time, show considerable promise in tracking the

dynamics of normal aging and dementia. The mean rate of brain

atrophy in AD, based on MRI measures of total cerebral volumes,

was recently reported to be 2.4 ± 1.1% per year in AD, compared

with 0.4 ± 0.5% per year in matched elderly controls (MMSE 19.6

± 4.1 and 29.2 ± 1.0 at baseline, for patients and controls,

respectively) (Fox et al., 2000). Higher rates of atrophy and

tissue loss have been estimated for specific structures, including

the hippocampus (Kaye et al., 1997; Jack et al., 1998; Laakso et

al., 2000). Four-dimensional maps of degenerative rates may also

be derived by computing a deformation field that elastically

transforms a subject’s anatomy from its earlier configuration to

its shape in a later scan (Fox et al., 1996, 2000; Thompson et al.,

2000b,d). We are currently extending the mapping approach

described here to store detailed population-based maps of

degenerative rates across time and explore linkages between

these maps and cognitive variables (Thompson et al., 2000d), as

well as therapeutic and genetic factors [e.g. ApoE genotype

(Small et al., 2000)].

Accurate mapping of gray matter changes in a living

population with AD holds significant promise for genetic,

longitudinal and interventional studies of dementia. In any study

where staging of the disease is required, the ability to calib-

rate gray matter integrity against a reference population is

paramount. The patient cohort on which our atlas is based is

being expanded to accommodate groups at different stages

of dementia. By following the same patients longitudinally

(Thompson et al., 2000b), statistical maps of gray matter loss at

multiple time points will ultimately provide a dynamic frame-

work to help understand the progression of the disease and to

gauge therapeutic, disease-modifying response in an individual

or clinically defined group.

Notes1. These additional boundaries included: (i) the posterior-medial limit of

the occipital lobe in each hemisphere, between the parieto-occipital

and posterior calcarine sulci; (ii) the inferior limit of the lingual

gyrus at the medial wall of each brain hemisphere, from the pos-

terior calcarine sulcus to the splenium of the corpus callosum; (iii) the

superior-medial boundary of the parietal lobe, from the parieto-

occipital to the central sulcus; (iv) the anterior boundary of the frontal

lobes, from the superior-medial limit of the central sulcus to the

antero-medial tip of the superior rostral sulcus; (v) the inferior

boundary of the frontal lobes, from the superior rostral sulcus

posteriorly and inferiorly along the rhinal gyri to the rostral tip of the

anterior commissure.

2. Significance levels. If there had been no pre-existing hypothesis on the

localization of significant gray matter loss, which was expected in the

temporal and temporo-parietal cortex, a correction for multiple

comparisons can be made. The significance threshold can be set at a

level derived from the effective number of resolution elements in the

statistical field (RESELs) (Worsley, 1994). This corrected P value

depends on the smoothness tensor of the residuals of the statistical

model, which can also be estimated from the surface data, using an

approach known as statistical f lattening (Worsley et al., 1999;

Thompson et al., 2000d).

This work was supported by a Human Brain Project grant to the

International Consortium for Brain Mapping, funded jointly by NIMH and

NIDA (P20 MH/DA52176), by a P41 Resource Grant from the NCRR

(RR13642), by NINCDS grant K08-NS01646, NIA grant K08-AG100784

and research grants from the National Library of Medicine

(LM/MH05639), the National Science Foundation (BIR 93-22434), the

NCRR (RR05956) and NINCDS/NIMH (NS38753).

Address correspondence to Paul Thompson, Room 4238, Reed

Neurological Research Center, Laboratory of Neuro Imaging, Department

of Neurology, UCLA School of Medicine, 710 Westwood Plaza, Los

Angeles, CA 90095–1769, USA. Email: [email protected]

ReferencesBilder RM, Wu H, Chakos MH, Bogerts B, Pollack S, Aronowitz J, Ashtari

M, Degreef G, Kane JM, Lieberman JA (1994) Cerebral morphometry

and clozapine treatment in schizophrenia. J Clin Psychiat 55 (suppl

B):53–56.

Brodmann K (1909) Vergleichende Lokalisationslehre der Grosshirnrinde

in ihren Prinzipien dargestellt auf Grund des Zellenbaues, Barth,

Leipzig. In: Some papers on the cerebral cortex. [Translated as:

Brodmann K (1960) On the comparative localization of the cortex, pp.

201–230. Springfield, IL: Thomas.]

Cao J, Worsley KJ (2000) The geometry of the Hotelling’s T-squared

random field with applications to the detection of shape changes. Ann

Statist (in press).

Clinton J, Blackman SEA, Royston MC, Robert GW (1994) Differential

synaptic loss in the cortex in AD: a study using archival material.

NeuroReport 5:497–500.

Collins DL, Le Goualher G, Venugopal R, Caramanos A, Evans AC, Barillot

C (1996) Cortical constraints for non-linear cortical registration, In:

Visualization in biomedical computing, Hamburg, Germany, Sept.

1996, Lecture notes in computer science (Höhne KH, Kikinis R, eds),

1131, pp. 307–316. Berlin: Springer Verlag.

Cook MJ, Free SL, Fish DR, Shorvon SD, Straughan K, Stevens JM (1994)

Analysis of cortical patterns, In: Magnetic resonance scanning and

epilepsy (Shorvon SD, ed), pp. 263–274. New York: Plenum.

Courchesne E (1997) Brainstem, cerebellar and limbic neuroanatomical

abnormalities in autism. Curr Opin Neurobiol 7:269–278.

Csernansky JG, Joshi S, Wang L, Haller JW, Gado M, Miller JP, Grenander

U, Miller MI (1998) Hippocampal morphometry in schizophrenia

by high dimensional brain mapping. Proc Natl Acad Sci USA

95:11406–11411.

Cuénod CA, Denys A, Michot JL, Jehenson P, Forette F, Kaplan D, Syrota

A, Boller F (1993) Amygdala atrophy in AD: an in vivo magnetic

resonance study. Arch Neurol 50:941–945.

Cummings J, Benson DF, LoVerme S (1980) Reversible dementia.

Illustrative cases, definition, and review. J Am Med Assoc 243:

2434–2439.

Cunningham DJ (1892) Contribution to the surface anatomy of the

cerebral hemispheres. Cunningham Memoirs (R. Irish Acad.) 7:372.

Davatzikos C (1996) Spatial normalization of 3D brain images using

deformable models. J. Comput. Assist Tomogr 20:656–665.

Davatzikos C, Vaillant M, Resnick SM, Prince JL, Letovsky S, Bryan RN

(1996) A computerized approach for morphological analysis of the

corpus callosum. J. Comput. Assist Tomogr 20:88–97.

Davidson RJ, Hugdahl K (1994) Brain asymmetry. Cambridge, MA: MIT

Press.

De La Monte SM (1989) Quantification of cerebral atrophy in pre-clinical

and end-stage AD. Ann Neurol 25:450–459.

Dinov ID,Mega MS, Thompson PM, Lee L, Woods RP, Holmes CJ, Sumners

DW, Toga AW (2000) Analyzing functional brain images in a

probabilistic atlas: a validation of sub-volume thresholding. J Comput

Assist Tomogr 24:128–138.

Drury HA, Van Essen DC (1997) Analysis of functional specialization in

human cerebral cortex using the Visible Man surface based atlas. Hum

Brain Map 5:233–237.

Drury HA, Van Essen DC, Joshi SC, Miller MI (1996) Analysis and

comparison of areal partitioning schemes using two-dimensional f luid

deformations. NeuroImage 3:S130.

Dumoulin SO, Bittar RG, Kabani NJ, Baker CL Jr, Le Goualher G, Pike B,

Evans AC (2000) A new anatomical landmark for reliable identification

of human area V5/MT: a quantitative analysis of sulcal patterning.

Cereb Cortex 10:454–463.

Cerebral Cortex Jan 2001, V 11 N 1 13

Page 14: Cortical Change in Alzheimer’s Disease Detected with a ...users.loni.ucla.edu/~thompson/CERCOR2001/cereb_ctx2001.pdf · specially constructed to reflect the average morphology of

Eberstaller O (1884) Zür Oberf lachen Anatomie der Grosshirn

Hemisphaeren. Wien Med Bl. 7:479,642,644.

Eidelberg D, Galaburda AM (1982) Symmetry and asymmetry in the

human posterior thalamus: I. Cytoarchitectonic analysis in normal

persons. Arch Neurol 39:325–332.

Evans AC, Collins DL, Neelin P, MacDonald D, Kamber M, Marrett TS

(1994) Three-dimensional correlative imaging: applications in human

brain mapping. In: Functional neuroimaging: technical foundations

(Thatcher RW, Hallett M, Zeffiro T, John ER, Huerta M, eds), pp.

145–162.

Falkai P, Bogerts B, Greve B, Pfeiffer U, Machus B, Folsch-Reetz B,

Majtenyi C, Ovary I (1992) Loss of Sylvian fissure asymmetry in

schizophrenia. A quantitative post mortem study. Schizophr Res

7:23–32.

Filipek PA, Kennedy DN, Caviness VS Jr, Rossnick SL, Spraggins TA,

Starewicz PM (1989) Magnetic resonance imaging-based brain

morphometry: development and application to normal subjects. Ann

Neurol 25:61–67.

Fischl B, Sereno MI, Tootell RBH, Dale AM (1999) High-resolution

inter-subject averaging and a coordinate system for the cortical

surface. Hum Brain Map 8:272–284.

Folstein MF, Folstein SE, McHugh PR (1975) ‘Mini mental state’: a

practical method of grading the cognitive state of patients for the

clinician. J Psychiat Res 12:189–198.

Fox NC, Freeborough PA, Rossor MN (1996) Visualisation and

quantification of rates of atrophy in AD. Lancet 348:94–97.

Fox NC, Cousens S, Scahill R, Harvey RJ, Rossor MN (2000) Using serial

registered brain magnetic resonance imaging to measure disease

progression in Alzheimer disease: power calculations and estimates of

sample size to detect treatment effects. Arch Neurol 57:339–344.

Fox PT, Huang AY, Parsons LM, Xiong JH, Rainey L, Lancaster JL (1999)

Functional volumes modeling: scaling for group size in averaged

images. Hum Brain Map 8:143–150.

Friedland RP, Luxenberg J (1988) Neuroimaging and dementia. In:

Clinical neuroimaging: frontiers in clinical neuroscience, Vol. 4

(Theodore WH, ed), pp. 139–163. New York: Allan Liss.

Galaburda AM (1995) Anatomic basis of cerebral dominance. In: Brain

asymmetry, (Davidson RJ, Hugdahl K, eds), pp. 51–73. Cambridge,

MA: MIT Press,.

Galaburda AM, Geschwind N (1981) Anatomical asymmetries in the adult

and developing brain and their implications for function. Adv Pediatr

28:271–292.

Ge Y, Fitzpatrick JM, Kessler RM, Jeske-Janicka M (1995) Intersubject

brain image registration using both cortical and subcortical

landmarks. SPIE Image Process 2434:81–95.

Gee JC, LeBriquer L, Barillot C, Haynor DR, Bajcsy R (1995) Bayesian

approach to the brain image matching problem, Institute for Research

in Cognitive Science Technical Report 95-08, April 1995.

Geschwind N, Levitsky W (1968) Human brain: Left-right asymmetries in

temporal speech region. Science 161:186.

Geyer S, Schormann T, Mohlberg H, Zilles K (2000) Areas 3a, 3b, and 1 of

human primary somatosensory cortex. NeuroImage 11:684–696.

Giedd JN, Castellanos FX, Casey BJ, Kozuch P, King AC, Hamburger SD,

Rapaport JL (1994) Quantitative morphology of the corpus callosum

in attention deficit hyperactivity disorder. Am J Psychiat

151:665–669.

Grenander U, Miller MI (1998) Computational anatomy: an emerging

discipline, Technical Report, Department of Mathematics, Brown

University.

Haller JW, Banerjee A, Christensen GE, Gado M, Joshi S, Miller MI,

Sheline Y, Vannier MW, Csernansky JG (1997) Three-dimensional

hippocampal MR morphometry with high-dimensional transfor-

mation of a neuroanatomic atlas. Radiology 202:504–510.

Holmes CJ, MacDonald D, Sled JG, Toga AW, Evans AC (1996) Cortical

peeling: CSF/grey/white matter boundaries visualized by nesting

isosurfaces. Proc Visualizat Biomed Comput 4:99–104.

Hubbard BM, Anderson JM (1981) A quantitative study of cerebral

atrophy in old age and senile dementia. J Neurol Sci 50:135–145.

Jack CR Jr, Petersen RC, Xu Y, O’Brien PC, Smith GE, Ivnik RJ, Tangalos

EG, Kokmen E (1998) Rate of medial temporal lobe atrophy in typical

aging and Alzheimer's disease. Neurology 51:993–999.

Jäncke L, Schlaug G, Huang Y, Steinmetz H (1994) Asymmetry of the

planum parietale. NeuroReport 5:1161–1163.

Jernigan TL, Salmon D, Butter N, et al. (1991) Cerebral structure on MRI,

Part II: specific changes in Alzheimer’s and Huntington’s diseases.

Biol Psychiat 29:68–81.

Johnson KA, Jones K, Holman BL, Becker JA, Spiers PA, Satlin A, Albert

MS (1998) Preclinical prediction of Alzheimer's disease using SPECT.

Neurology 50:1563–1571.

Joshi S, Miller MI, Grenander U (1998) On the geometry and shape of

brain sub-manifolds. Int J Pattern Recogn Artif Intell 11:1317–1343.

Kaye JA, Swihart T, Howieson D, Dame A, Moore MM, Karnos T, Camicioli

R, Ball M, Oken B, Sexton G (1997) Volume loss of the hippocampus

and temporal lobe in healthy elderly persons destined to develop

dementia. Neurology 48:1297–1304.

Kennedy DN, Lange N, Makris N, Bates J, Meyer J, Caviness VS Jr (1998)

Gyri of the human neocortex: an MRI-based analysis of volume and

variance. Cereb Cortex 8:372–384.

Kikinis R, Shenton ME, Gerig G, Hokama H, Haimson J, O’Donnell BF,

Wible CG, McCarley RW, Jolesz FA (1994) Temporal lobe sulco-gyral

pattern anomalies in schizophrenia: an in vivo MR three-dimensional

surface rendering study. Neuroscience Lett 182:7–12.

Kulynych JJ, Vladar K, Jones DW, Weinberger DR (1996) Superior

temporal gyrus volume in schizophrenia: a study using MRI

morphometry assisted by surface rendering. Am J Psychiat 153:50–56.

Laakso MP, Lehtovirta M, Partanen K, Riekkinen PJ, Soininen H (2000)

Hippocampus in AD: a 3-year follow-up MRI study. Biol Psychiat

47:557–561.

Le Goualher G, Barillot C, Bizais Y, Scarabin J-M, (1996) 3D segmentation

of cortical sulci using active models. SPIE Med Imaging 2710,

254–263.

Leonard CM (1996) Structural variation in the developing and mature

cerebral cortex: noise or signal? In: Developmental neuroimaging:

mapping the development of brain and behavior (Thatcher RW, Reid

Lyon G, Rumsey J, Krasnegor N, eds) pp. 207–231. Academic Press.

Loewenstein DA, Barker WW, Chang JY, Apicella A, Yoshii F, Kothari P,

Levin B, Duara R (1989) Predominant left hemisphere metabolic

dysfunction in dementia. Arch. Neurol 46:146–152.

MacDonald D (1998) A method for identifying geometrically simple

surfaces from three dimensional images, PhD Thesis, McGill

University, Canada.

MacDonald D, Avis D, Evans AC (1993) Automatic parameterization of

human cortical surfaces. In Annual Symposium on Information

Processing in Medical Imaging (IPMI).

MacDonald D, Avis D, Evans AC (1994) Multiple surface identification and

matching in magnetic resonance imaging. Proc SPIE 2359:160–169.

MacDonald D, Venugopal R, Caramanos Z, Petrides M, Avis D, Evans AC

(1997) A surface-based 2D sulcal atlas. NeuroImage 5:S414.

Malobabic S, Marinkovic R, Lesic A, Draganic S, Duranovic S, Sojic M

(1993) Morphologic asymmetry of the frontal lobe of the cerebral

hemisphere in man. Med Pregl 46:401–405.

Manceaux-Demiau A, Bryan RN, Davatzikos C (1998) A probabilistic

ribbon model for shape analysis of the cerebral sulci: application to

the central sulcus. J Comput Assist Tomogr 22:962–971.

Mangin J-F, Frouin V, Bloch I, Regis J, López–Krahe J (1995) From 3D

magnetic resonance images to structural representations of the cortex

topography using topology-preserving deformations. J Math Imaging

Vision 5: 297–318.

Mazziotta JC, Toga AW, Evans AC, Fox P, Lancaster J (1995) A probabilistic

atlas of the human brain: theory and rationale for its development.

NeuroImage 2:89–101.

McKhann G, Drachman D, Folstein M, Katzman R, Price D, Stadian EM

(1984) Clinical diagnosis of AD: report of the NINCDS-ARDRA Work

Group under the Auspices of the Health and Human Services Task

Force on Alzheimer’s Disease. Neurology 34:939–944.

Mega MS, Chen S, Thompson PM, Woods RP, Karaca TJ, Tiwari A, Vinters

H, Small GW, Toga AW (1997a) Mapping pathology to metabolism:

coregistration of stained whole brain sections to PET in AD.

NeuroImage 5:147–153.

Mega MS, Chu T, Thompson PM, Mazziotta JC, Burt J, Aron J, Ghasri P,

Chen S, Lim J, Cole GM, Toga AW (1997b) [18F]-Fluorodeoxyglucose

positron emission tomography (FDG-PET) corrected with synapto-

physin density is inversely related to beta-amyloid burden in AD. Ann

Neurol 42:M23.

Mega MS, Thompson PM, Cummings JL, Back CL, Xu LQ, Zohoori S,

Goldkorn A, Moussai J, Fairbanks L, Small GW, Toga AW (1998) Sulcal

variability in the Alzheimer’s brain: correlations with cognition.

Neurology, 50:145–151.

Mega MS, Thompson PM, Toga AW, Cummings JL (2000a) Brain mapping

14 Cortical Change in Alzheimer’s Disease • Thompson et al.

Page 15: Cortical Change in Alzheimer’s Disease Detected with a ...users.loni.ucla.edu/~thompson/CERCOR2001/cereb_ctx2001.pdf · specially constructed to reflect the average morphology of

in dementia. In: Brain mapping: the disorders (Toga AW, Mazziotta JC,

eds). Academic Press (in press).

Mega MS, Thompson PM, Dinov ID, Toga AW, Cummings JL (2000b) The

UCLA Alzheimer Brain Atlas Project: structural and functional

applications. In: Proceedings of the 2000 World Alzheimer’s

Congress, Washington, DC.

Missir O, Dutheil-Desclercs C, Meder JF, Musolino A, Fredy D (1989)

Central sulcus patterns at MRI. J Neuroradiol 16:133–144.

Murphy DGM, DeCarli CD, Daly E, Gillette JA, McIntosh AR, Haxby JV,

Teichberg D, Schapiro MB, Rapoport SI, Horwitz B (1993) Volumetric

magnetic resonance imaging in men with dementia of the Alzheimer

type: correlations with disease severity. Biol Psychiat 34:612–621.

Novikov II, Podcherednik TN (1992) Variability of the gyri and sulci of the

temporal lobe of the human brain. Zh Nevropatol Psikhiatr S S

Korsakova 92:102–105.

Ono M, Kubik S, Abernathey CD (1990) Atlas of the cerebral sulci.

Stuttgart: Thieme.

Paus T, Tomaioulo F, Otaky N, MacDonald D, Petrides M, Atlas J, Morris R,

Evans AC (1996) Human cingulate and paracingulate sulci: pattern,

variability, asymmetry and probabilistic map. Cereb Cortex

6:207–214.

Pitiot A, Thompson PM, Toga AW (2000) Spatially and temporally

adaptive elastic template matching. IEEE Trans Pattern Anal Machine

Intell (in press).

Rademacher J, Caviness VS Jr, Steinmetz H, Galaburda AM (1993)

Topographical variation of the human primary cortices: implications

for neuroimaging, brain mapping, and neurobiology. Cereb Cortex

3:313–329.

Rajkowska G, Goldman-Rakic P (1995) Cytoarchitectonic definition of

pre-frontal areas in the normal human cortex: II. Variability in loca-

tions of areas 9 and 46 and relationship to the Talairach coordinate

system. Cereb Cortex 5:323–337.

Roberts GW, Leigh PN, Weinberger DR (1993) Neuropsychiatric

disorders, Ch 2(1). Gower Medical Publishers.

Royackkers N, Desvignes M, Revenu M (1996) Construction automatique

d’un atlas adaptatif des sillons corticaux, ORASIS 96, Clermont-

Ferrand, pp. 187–192.

Sled JG, Zijdenbos AP, Evans AC (1998) A non-parametric method for

automatic correction of intensity non-uniformity in MRI data. IEEE

Trans Med Imaging 17:87–97.

Sobire G, Goutieres F, Tardieu M, Landrieu P, Aicardi J (1995) Extensive

macrogyri or no visible gyri: distinct clinical, electroencephalo-

graphic, and genetic features according to different imaging patterns.

Neurology 45:1105–1111.

Sowell ER, Thompson PM, Holmes CJ, Batth R, Trauner DA, Jernigan TL,

Toga AW (1999a) Localizing age-related changes in brain structure

between childhood and adolescence using statistical parametric

mapping. NeuroImage 9:587–597.

Sowell ER, Thompson PM, Holmes CJ, Jernigan TL, Toga AW (1999b)

Progression of structural changes in the human brain during the first

three decades of life: in vivo evidence for post-adolescent frontal and

striatal maturation. Nature Neurosci 2:859–861.

Steinmetz H, Furst G, Freund H-J (1989) Cerebral cortical localization:

application and validation of the proportional grid system in MR

imaging. J Comput Assist Tomogr 13:10–19.

Steinmetz H, Furst G, Freund H-J (1990) Variation of perisylvian and

calcarine anatomic landmarks within stereotaxic proportional

coordinates. Am J Neuroradiol 11:1123–1130.

Strauss E, Kosaka B, Wada J (1983) The neurobiological basis of lateralized

cerebral function. A review. Hum Neurobiol 2:115–127.

Talairach J, Tournoux P (1988) Co-planar stereotaxic atlas of the human

brain. New York: Thieme.

Thirion J-P, Prima S, Subsol S (1998) Statistical analysis of dissymmetry in

volumetric medical images. Med Image Anal (in press).

Thompson PM, Toga AW (1997) Detection, visualization and animation of

abnormal anatomic structure with a deformable probabilistic brain

atlas based on random vector field transformations. Med Image Anal

1:271–294.

Thompson PM, Toga AW (1998) Anatomically-driven strategies for

high-dimensional brain image warping and pathology detection. In:

Brain warping (Toga AW, ed), pp. 311–336. Academic Press.

Thompson PM, Schwartz C, Toga AW (1996a) High-resolution random

mesh algorithms for creating a probabilistic 3D surface atlas of the

human brain. NeuroImage 3:19–34.

Thompson PM, Schwartz C, Lin RT, Khan AA, Toga AW (1996b) 3D

statistical analysis of sulcal variability in the human brain. J Neurosci

16:4261–4274.

Thompson PM, MacDonald D, Mega MS, Holmes CJ, Evans AC, Toga AW

(1997) Detection and mapping of abnormal brain structure with a

probabilistic atlas of cortical surfaces. J Comp Assist Tomogr

21:567–581.

Thompson PM, Moussai J, Khan AA, Zohoori S, Goldkorn A, Mega MS,

Small GW, Cummings JL, Toga AW (1998) Cortical variability and

asymmetry in normal aging and Alzheimer’s Disease. Cereb Cortex

8:492–509.

Thompson PM, Woods RP, Mega MS, Toga AW (2000a) Mathematical/

computational challenges in creating population-based brain atlases.

Hum Brain Map 9:81–92.

Thompson PM, Giedd JN, Woods RP, MacDonald D, Evans AC, Toga AW

(2000b) Growth patterns in the developing human brain detected

using continuum-mechanical tensor mapping. Nature, 404:190–193.

Thompson PM, Mega MS, Cummings JL, Toga AW (2000c) Detecting

dynamic (4D) profiles of degenerative rates in AD patients, using

tensor mapping and a population-based brain atlas. Proc Soc

Neurosci.

Thompson PM, Mega MS, Narr KL, Sowell ER, Blanton RE, Toga AW

(2000d) Brain image analysis and atlas construction. In: SPIE

Handbook on Medical Image Analysis (Fitzpatrick M, ed). Society of

Photo-Optical Instrumentation Engineers (SPIE) Press.

Toga AW, Thompson PM (1998) Multimodal brain atlases. In: Advances

in biomedical image databases (Wong S, ed), pp.53–88. Kluwer

Academic Press.

Van Essen DC, Drury HA (1997) Structural and functional analyses

of human cerebral cortex using a surface-based atlas. J Neurosci

17:7079–7102.

Watson JD, Myers R, Frackowiak RS, Hajnal JV, Woods RP, Mazziotta JC,

Shipp S, Zeki S (1993) Area V5 of the human brain: evidence from a

combined study using positron emission tomography and magnetic

resonance imaging. Cereb Cortex 3:79–94.

Witelson SF, Kigar DL (1992) Sylvian fissure morphology and asymmetry

in men and women: bilateral differences in relation to handedness in

men. J Comp Neurol 323:326–340.

Woods RP, Mazziotta JC, Cherry SR (1993) MRI-PET pegistration with

automated algorithm. J Comput Assist Tomogr 17:536–546.

Woods RP, Grafton ST, Watson JDG, Sicotte NL, Mazziotta JC. (1998)

Automated image registration: II. Intersubject validation of linear and

nonlinear models. J Comput Assist Tomogr 22:153–165.

Worsley KJ (1994) Local maxima and the expected Euler characteristic of

excursion sets of chi-squared, F and t fields. Adv Appl Probab

26:13–42.

Worsley KJ, Andermann M, Koulis T, MacDonald D, Evans AC (1999)

Detecting changes in non-isotropic images. Hum Brain Map 8:98–101.

Zeineh MM, Thompson PM, Engel SA, Bookheimer SY (2000) Averaging

f lat maps of hippocampal activity across subjects, 6th International

Conference on Functional Mapping of the Human Brain, San Antonio,

Texas, June 2000.

Zijdenbos AP, Dawant BM (1994) Brain segmentation and white matter

lesion detection in MR images. Crit Rev Biomed Eng 22:401–465.

Appendix

Constructing Variability Maps

The imposition of a standard surface grid on each subject’s cortex makes

it easier to compare anatomical models from multiple subjects. By

averaging nodes with the same grid coordinates across subjects, an

average surface is produced for each group. The mean surface acts as a

reference surface, relative to which deviations (displacements) in the

other surfaces are measured. Information on subjects’ individual

differences is then stored as a vector-valued displacement map, indicating

how a subject deviates locally from the average anatomy. Color maps that

illustrate the spatial variability of anatomy in a group were computed as

by Thompson and co-workers (Thompson et al., 1996b). Brief ly, for a

group of n subjects, the variability in spatial position for points ri(u,v)

internal to a particular anatomical surface is computed, based on the

maps, as a scalar variance function:

σ2(u,v) = (1/[n – 1])Σi = 1 to n||ri(u,v) – rµ(u,v)||2 (1)

defined at each mesh node (u,v), where rµ(u,v) is the average surface.

Cerebral Cortex Jan 2001, V 11 N 1 15

Page 16: Cortical Change in Alzheimer’s Disease Detected with a ...users.loni.ucla.edu/~thompson/CERCOR2001/cereb_ctx2001.pdf · specially constructed to reflect the average morphology of

The square root of this function gives the standard deviation in

stereotaxic position as a 3D r.m.s. distance for each internal surface point.

The appropriate numerical value, at each grid point, is given by the root

mean square magnitude of the 3D displacement vectors assigned to that

point, in the n surface maps from the individual to average. The variability

measure is visualized using a color code to illustrate the profile of

variability across the anatomy.

Elastic Matching of Gyral Patterns using Flow Fields

Differences in cortical patterns between any pair of subjects were

determined by deforming one cortical model to match the other. This

procedure has been covered in detail by Thompson and co-workers

(Thompson et al., 2000a) and is summarized here for completeness. Since

each cortical model is obtained by deforming a spherical surface into

the shape of the cortex, gyral features can be mapped back onto a sphere

and, subsequently, to a plane (Fig. 2). This simplifies computation of

anatomical correspondences. Anatomical correspondences can therefore

be computed by defining a f low field in the f lat, 2D parameter space that

matches gyral features from one subject to another (Figs 2, 3) (Davatzikos

et al., 1996; Thompson et al., 1996a, 1997, 2000d; Fischl et al., 1999).

The f low is given by the solution to a curve-driven warp in the f lat

parametric space of the cortex (Thompson et al., 1996, 1998, 2000). The

f low behavior is modeled using equations derived from continuum

mechanics and these equations are governed by the Cauchy–Navier

differential operator L = µ∇ 2 + (λ + µ)∇ (∇ T) (Davatzikos et al., 1996;

Thompson et al., 1996, 1998, 2000d; Grenander and Miller, 1998).

Technical Details

Specifically, for points r = (r,s) in the cortical parameter space Ω = [0,2π)

× [0,π), a system of simultaneous partial differential equations can be

written for the f low field u (r):

L‡(u(r)) + F(r – u(r)) = 0, ∀ r ∈ Ω ,

with u(r) = u0(r), ∀ r ∈ M0∪ M1

(2)

Here M0, M1 are sets of points and (sulcal or gyral) curves where

displacement vectors u(r) = u0(r) matching the corresponding anatomy

across subjects are known. The f low behavior is governed by the

Cauchy–Navier differential operator L = µ∇ 2 + (λ + µ)∇ (∇ T) with body

force F (Thompson et al., 1996, 1998, 2000; Grenander and Miller, 1998).

In solving this governing equation matching sulcal networks across

subjects, dependencies between the metric tensors of the surface para-

meterizations and the matching field are eliminated with an approach

known as covariant regularization, which uses generalized coordinates

and correction terms known as Christoffel symbols (Thompson and Toga,

2000a,d). Because of the intrinsic curvature of the cortex, this means that

the ‘covariant form’ L‡ of the differential operator L is used when solving

these equations (Thompson and Toga, 1998; Thompson et al., 2000d).

This adjustment also makes sure that cortical surfaces are matched in a

way that is actually independent of the way the surfaces are f lattened;

in other words, the matching procedure is parameterization-invariant. In

the partial differential equations (2) we replace L by the covariant

differential operator L‡. In L‡ all L value partial derivatives are replaced by

covariant derivatives. These covariant derivatives are defined with respect

to the metric tensor of the surface domain where calculations are

performed. The covariant derivative of a (contravariant) vector field,

ui(x), is defined as ui,k = ∂uj/∂xk + Γj

ik ui, where the Christoffel symbols of

the second kind, Γjik, are computed from derivatives of the metric tensor

components gjk(x):

Γjik = (1/2) gil (∂glj/∂xk + ∂glk/∂xj – ∂gjk/∂xi) (3)

These correction terms are then used in the elastic transformation used to

match one cortex with another.

Finally, because 3D cortical positions are encoded in color on the f lat

maps, the surface matching transformation is recovered in 3D as a

mapping that drives one cortex onto another. A color code (Fig. 2e)

representing 3D cortical point locations in an individual subject is

convected along with the f low that drives the sulcal pattern into the

average configuration for the group (Fig. 2f). Once this is done in all

subjects at a particular location in the f lat map (Fig. 2f), points on each

individual’s cortex are recovered that have the same relative location

to the primary folding pattern in all subjects. Averaging of these

corresponding points results in a crisp average cortex (Fig. 3, bottom

row). The corresponding 3D displacement is recovered between the

cortical models. This displacement matches a large network of sulcal

features and thus is a valid encoding of gyral pattern differences.

16 Cortical Change in Alzheimer’s Disease • Thompson et al.


Recommended