ArticlePDF Available

Brain health in diverse settings: How age, demographics and cognition shape brain function

Authors:
NeuroImage 295 (2024) 120636
Available online 21 May 2024
1053-8119/© 2024 The Author(s). Published by Elsevier Inc. This is an open access article under the CC BY license (http://creativecommons.org/licenses/by/4.0/).
Brain health in diverse settings: How age, demographics and cognition
shape brain function
Hernan Hernandez
a
,
1
, Sandra Baez
b
,
c
,
1
, Vicente Medel
a
, Sebastian Moguilner
a
,
d
,
Jhosmary Cuadros
a
,
e
,
f
, Hernando Santamaria-Garcia
g
,
h
, Enzo Tagliazucchi
a
,
i
, Pedro A. Valdes-
Sosa
k
,
l
, Francisco Lopera
m
, John Fredy OchoaG´
omez
m
, Alfredis Gonz´
alez-Hern´
andez
n
,
Jasmin Bonilla-Santos
o
, Rodrigo A. Gonzalez-Montealegre
p
, Tuba Aktürk
ap
, Ebru Yıldırım
ap
,
Renato Anghinah
q
,
r
, Agustina Legaz
a
,
s
,
t
,
u
, Sol Fittipaldi
a
,
c
,
ab
, G¨
orsev G. Yener
v
,
w
,
x
,
Javier Escudero
y
, Claudio Babiloni
z
,
aa
, Susanna Lopez
z
, Robert Whelan
c
,
ac
, Alberto A
Fern´
andez Lucas
ac
, Adolfo M. García
c
,
ad
,
ae
, David Huepe
af
, Gaetano Di Caterina
ag
, Marcio Soto-
A˜
nari
ah
, Agustina Birba
a
, Agustin Sainz-Ballesteros
a
, Carlos Coronel
a
,
c
,
ai
, Eduar Herrera
aj
,
Daniel Abasolo
ak
, Kerry Kilborn
al
, Nicol´
as Rubido
am
, Ruaridh Clark
an
, Ruben Herzog
a
,
ao
,
Deniz Yerlikaya
ap
, Bahar Güntekin
aq
,
ar
, Mario A. Parra
as
, Pavel Prado
j
,
Agustin Ibanez
a
,
at
,
au
,
av
,
*
a
Latin American Brain Health Institute, Universidad Adolfo Iba˜
nez, Santiago de Chile, Chile
b
Universidad de los Andes, Bogota, Colombia
c
Global Brain Health Institute (GBHI), University of California, San Francisco, US Trinity College Dublin, Dublin, Ireland
d
Harvard Medical School, Boston, MA, USA
e
Advanced Center for Electrical and Electronic Engineering, Universidad T´
ecnica Federico Santa María, Valparaíso 2390123, Chile
f
Grupo de Bioingeniería, Decanato de Investigaci´
on, Universidad Nacional Experimental del T´
achira, San Crist´
obal 5001, Venezuela
g
Ponticia Universidad Javeriana (PhD Program in Neuroscience) Bogot´
a, San Ignacio, Colombia
h
Center of Memory and Cognition Intellectus, Hospital Universitario San Ignacio Bogot´
a, San Ignacio, Colombia
i
University of Buenos Aires, Argentina
j
Escuela de Fonoaudiología, Facultad de Odontología y Ciencias de la Rehabilitaci´
on, Universidad San Sebasti´
an, Santiago, Chile
k
The Clinical Hospital of Chengdu Brain Sciences, University of Electronic Sciences Technology of China, Chengdu, China
l
Cuban Neuroscience Center, La Habana, Cuba
m
Grupo de Neurociencias de Antioquia, University of Antioquia, Medellín, Colombia
n
Master programme of Clinical Neuropsychology, Universidad Surcolombiana, Neiva Huila, Colombia
o
Department of Psychology, Universidad Cooperativa de Colombia, Colombia
p
Neurocognition and Psychophysiology Laboratory, Universidad Surcolombiana, Neiva Huila, Colombia
q
Reference Center of Behavioural Disturbances and Dementia, School of Medicine, University of Sao Paulo, Sao Paulo, Brazil
r
Traumatic Brain Injury Cognitive Rehabilitation Out-Patient Center, University of Sao Paulo, Sao Paulo, Brazil
s
Cognitive Neuroscience Center, Universidad de San Andr´
es, Buenos Aires, Argentina
t
National Scientic and Technical Research Council (CONICET), Buenos Aires, Argentina
u
Facultad de Psicología, Universidad Nacional de C´
ordoba, C´
ordoba, Argentina
v
Faculty of Medicine, Izmir University of Economics, 35330, Izmir, Turkey
w
Brain Dynamics Multidisciplinary Research Center, Dokuz Eylul University, Izmir, Turkey
x
Izmir Biomedicine and Genome Center, Izmir, Turkey
y
School of Engineering, Institute for Imaging, Data and Communications, University of Edinburgh, Scotland, UK
z
Department of Physiology and Pharmacology V. Erspamer, Sapienza University of Rome, Rome, Italy
aa
Hospital San Raffaele Cassino, Cassino, (FR), Italy
ab
School of Psychology, Trinity College Dublin, Dublin, Ireland
ac
Department of Legal Medicine, Psychiatry and Pathology at the Complutense University of Madrid, Madrid, Spain
ad
Cognitive Neuroscience Center, Universidad de San Andr´
ess, Buenos Aires, Argentina
ae
Departamento de Lingüística y Literatura, Facultad de Humanidades, Universidad de Santiago de Chile, Santiago, Chile
af
Center for Social and Cognitive Neuroscience (CSCN), School of Psychology, Universidad Adolfo Ib´
a˜
nez
ag
Department of Electronic and Electrical Engineering, University of Strathclyde, Glasgow, UK
ah
Universidad Cat´
olica San Pablo, Arequipa, Peru
* Corresponding author.
E-mail address: agustin.ibanez@gbhi.org (A. Ibanez).
Contents lists available at ScienceDirect
NeuroImage
journal homepage: www.elsevier.com/locate/ynimg
https://doi.org/10.1016/j.neuroimage.2024.120636
Received 8 February 2024; Received in revised form 17 April 2024; Accepted 4 May 2024
NeuroImage 295 (2024) 120636
2
ai
Centro Interdisciplinario de Neurociencia de Valparaíso (CINV), Universidad de Valparaíso, Valparaíso, Chile
aj
Departamento de Estudios Psicol´
ogicos, Universidad ICESI, Cali, Colombia
ak
Centre for Biomedical Engineering, School of Mechanical Engineering Sciences, Faculty of Engineering and Physical Sciences, University of Surrey, Guildford GU2 7XH,
UK
al
School of Psychology, University of Glasgow, Glasgow, Scotland, UK
am
Institute for Complex Systems and Mathematical Biology, University of Aberdeen, Aberdeen, AB24 3UE, UK
an
Centre for Signal and Image Processing, Department of Electronic and Electrical Engineering, University of Strathclyde, UK
ao
Sorbonne Universit´
e, Institut du Cerveau - Paris Brain Institute - ICM, Inserm, CNRS, Paris 75013, France
ap
Department of Neurosciences, Health Sciences Institute, Dokuz Eylül University, Izmir, Turkey
aq
Health Sciences and Technology Research Institute (SABITA), Istanbul Medipol University, Istanbul, Turkey
ar
Department of Biophysics, School of Medicine, Istanbul Medipol University, Turkey
as
Department of Psychological Sciences and Health, University of Strathclyde, United Kingdom and Associate Researcher of the Latin American Brain Health Institute
(BrainLat), Universidad Adolfo Ib´
a˜
nez, Santiago, Chile
at
Global Brain Health Institute, University of California San Francisco, San Francisco, CA, USA
au
Cognitive Neuroscience Center, Universidad de San Andr´
es and Consejo Nacional de Investigaciones Cientícas y T´
ecnicas, Buenos Aires, Argentina
av
Trinity College Dublin, The University of Dublin, Dublin, Ireland
ARTICLE INFO
Keywords:
Age
Cognition
Education
Individual differences
Sex
Brain dynamics
ABSTRACT
Diversity in brain health is inuenced by individual differences in demographics and cognition. However, most
studies on brain health and diseases have typically controlled for these factors rather than explored their po-
tential to predict brain signals. Here, we assessed the role of individual differences in demographics (age, sex,
and education; n =1298) and cognition (n =725) as predictors of different metrics usually used in case-control
studies. These included power spectrum and aperiodic (1/f slope, knee, offset) metrics, as well as complexity
(fractal dimension estimation, permutation entropy, Wiener entropy, spectral structure variability) and con-
nectivity (graph-theoretic mutual information, conditional mutual information, organizational information) from
the source space resting-state EEG activity in a diverse sample from the global south and north populations.
Brain-phenotype models were computed using EEG metrics reecting local activity (power spectrum and
aperiodic components) and brain dynamics and interactions (complexity and graph-theoretic measures). Elec-
trophysiological brain dynamics were modulated by individual differences despite the varied methods of data
acquisition and assessments across multiple centers, indicating that results were unlikely to be accounted for by
methodological discrepancies. Variations in brain signals were mainly inuenced by age and cognition, while
education and sex exhibited less importance. Power spectrum activity and graph-theoretic measures were the
most sensitive in capturing individual differences. Older age, poorer cognition, and being male were associated
with reduced alpha power, whereas older age and less education were associated with reduced network inte-
gration and segregation. Findings suggest that basic individual differences impact core metrics of brain function
that are used in standard case-control studies. Considering individual variability and diversity in global settings
would contribute to a more tailored understanding of brain function.
1. Introduction
Individual differences encompass variations observed within a pop-
ulation in relation to a specic trait, including socio-demographic
characteristics, as well as psychological and cognitive factors (Senner
et al., 1814). Individual differences in age (Cole et al., 2018; Bethlehem
et al., 2022), sex (Dotson and Duarte, 2020; Tang et al., 2022; Allouh
et al., 2020), education (Garcia et al., 2018; Members et al., 2010), and
cognition (Valsd´
ottir et al., 2022; Heger et al., 2021) are currently
assumed to play a relevant role in brain function, driving diversity in
brain health (Santamaria-Garcia et al., 2023; Greene et al., 2022; Aranda
et al., 2021; Livingston et al., 2020; Holmes and Patrick, 2018). Di-
versity in the context of brain health refers to wide range of factors
related to populations in terms of race, ethnicity, gender, age, socio-
economic status, cognitive variability, geographical location, genetic
background, and health status, among others (Santamaria-Garcia et al.,
2023; Fittipaldi et al., 2023; Matshabane, 2021). Diversity impacts
healthy aging (Santamaria-Garcia et al., 2023), psychiatric conditions
(Holmes and Patrick, 2018; Resende et al., 2019), and neuro-
degeneration (Ibanez et al., 2023; Walsh et al., 2022). However, these
effects are not universally shared across regions (Santamaria-Garcia
et al., 2023; Greene et al., 2022; Holmes and Patrick, 2018; Alladi and
Hachinski, 2018), and may not generalize to more diverse, underrep-
resented populations in brain health research (Dotson and Duarte, 2020;
Santamaria-Garcia et al., 2023; Baez et al., 2023).
Recent ndings underscore the signicance of individual differences
in demographic predictors, showing that brain-phenotype models pri-
marily reect these variables rather than the intended cognitive do-
mains, particularly in non-stereotypical populations (Greene et al.,
2022). Notwithstanding, most studies on brain health and aging
(Jawinski et al., 2022; Yan et al., 2015), psychiatry (Wolfers et al., 2020;
Elad et al., 2021), and neurodegeneration (Pinaya et al., 2021; Verdi
et al., 2022) have typically controlled for these variables rather than
investigating their impact on brain-phenotype associations. Incorpo-
rating the role of individual differences, including demographics and
cognition, in functional brain dynamics across large, heterogeneous and
underrepresented populations may provide a direct pathway to a better
understanding of the impact of diversity on brain health.
The study of demographics and cognition as predictors of brain
signals has predominantly relied on MRI or fMRI (Bethlehem et al.,
2022; Elad et al., 2021; Di Biase et al., 2023; Rutherford et al., 2022;
Habes et al., 2021; Rosenberg et al., 2020). A limited number of reports
have incorporated MEG (Dimitriadis, 2022; Stier et al., 2023), and only
one has used EEG (Hill et al., 2022). Age has consistently emerged as the
most frequently studied demographic factor for predicting structural
and functional changes in the human brain, spanning from
post-conception to aging, both in healthy individuals (Bethlehem et al.,
2022; Di Biase et al., 2023; Rutherford et al., 2022; Dimitriadis, 2022;
Stier et al., 2023) and in patients with psychiatric (Elad et al., 2021) or
neurodegenerative conditions (Bethlehem et al., 2022). While sex is
considered a predictor of brain changes across the lifespan (Bethlehem
et al., 2022; Rutherford et al., 2022; Stier et al., 2023), the impacts of
other relevant variables, such as education and cognition (Rosenberg
1
Equal contribution.
H. Hernandez et al.
NeuroImage 295 (2024) 120636
3
et al., 2020; Greene et al., 2020; Garo-Pascual et al., 2023; Cesnaite
et al., 2023), have been less explored. Both education (Gordon et al.,
2008; Jokinen et al., 2016) and cognition (Habes et al., 2021) have been
associated with white matter changes and increased functional con-
nectivity (Franzmeier et al., 2071a, 2017b). However, while previous
studies have investigated how demographic and cognitive factors in-
uence brain signals, few have explored the effects of age, sex, educa-
tion, and cognition simultaneously. This research gap highlights the
need for studies that comprehensively explore the interplay between
demographic factors, cognitive abilities, and brain signals using various
imaging modalities, including EEG. These advancements would repre-
sent a crucial step towards robustly quantifying individual variations
and their associations with signatures of brain functional organization or
reorganization, especially in global research settings exhibiting larger
heterogeneity.
Given the limited accessibility of neuroimaging or MEG measures on
a large scale, more cost-effective and scalable techniques, such as EEG,
may be better suited for studying individual differences in global set-
tings. Neuroimaging measures often exhibit poor reliability and low
power, requiring thousands of participants to demonstrate brain-
phenotype associations (Marek et al., 2022). The non-invasiveness,
scalability, availability, temporal resolution, and low cost of EEG
make it an attractive approach to meet this need (Prado et al., 2022).
Although many studies have explored the association of EEG metrics
with demographics and cognition, only one study have used these var-
iables to predict EEG signals (Hill et al., 2022). Studies have identied
age-related EEG changes in resting and task performance (Valsd´
ottir
et al., 2022; Cesnaite et al., 2023; Al Zoubi et al., 2018; Carrier et al.,
2001; Hinault et al., 2023; Javaid et al., 2022; Merkin et al., 2023;
Meunier et al., 2009; Murty et al., 2020; Smith et al., 2023; Smits et al.,
2016; Song et al., 2014; Stacey et al., 2021; Trammell et al., 2017;
Trondle et al., 2023; Valdes-Sosa et al., 2021; Voytek et al., 2015;
Zappasodi et al., 2015), including alpha amplitude, as well as slowdown
and increases in global power (Cesnaite et al., 2023; Trondle et al., 2023;
Gaal et al., 2010), topographic reorganization in delta and theta fre-
quencies (Ishii et al., 2017; Rossini et al., 2007), and attenuated spon-
taneous gamma oscillations (Murty et al., 2020). Complexity changes
across the lifespan with increases during young adulthood and decreases
in the elderly population (Zappasodi et al., 2015) have been reported.
Age also induces changes in the aperiodic components (Hinault et al.,
2023; Trondle et al., 2023; Donoghue et al., 2020) and reductions in
global and local efciencies as well as small-worldness (Javaid et al.,
2022; Meunier et al., 2009; Achard and Bullmore, 2007; Onoda and
Yamaguchi, 2013). The association of sex and EEG metrics is less
conclusive (Carrier et al., 2001; Pravitha et al., 2005; Miraglia et al.,
2015). Conversely, the link between cognition and EEG signals is
well-established, with theta (Finnigan and Robertson, 2011) and alpha
(Lejko et al., 2020) associations, decrease in random and spontaneous
neural activity (Smith et al., 2023; Ouyang et al., 2020; Pei et al., 2023),
increased complexity (Smith et al., 2023; Ouyang et al., 2020; Pei et al.,
2023), higher network integration (Finnigan and Robertson, 2011;
Bullmore and Sporns, 2009; Iinuma et al., 2022; McBride et al., 2014),
and small-world properties (Liao et al., 2017). The association between
education and EEG metrics has been rarely reported, with limited evi-
dence (Wilkinson et al., 2023). Furthermore, although EEG studies have
improved preprocessing pipelines (Prado et al., 2022; Bigdely-Shamlo
et al., 2015; Ballesteros et al., 2023; Prado et al., 2023), several meth-
odological caveats continue to hinder their effective use in large-scale
studies. The recording heterogeneity, different layouts of electrodes
and ampliers, lack of harmonization, disparate processing pipelines,
and small sample sizes, have restricted broader application of EEG
(Prado et al., 2022; Jovicich et al., 2019). Thus, there is a critical need
for studies in large and diverse samples using demographic or cognitive
variables to predict brain signals via EEG while addressing its sources of
heterogeneity.
To address these gaps, this study explored EEG data from a large and
diverse sample of healthy participants under resting-state conditions
from regions in the global south (Argentina, Brazil, Colombia, Chile,
Cuba) and the global north (Ireland, Italy, Turkey, United Kingdom). We
aimed to validate the relationships between demographic (i.e., age, sex,
and education; n =1298) and cognitive factors (n =725) with the EEG
metrics across a diverse, heterogeneous, and global dataset. Addition-
ally, we investigated their utility in predicting signicant changes in
brain function, even in the presence of large heterogeneity. We
computed brain-phenotype models of local activity [i.e., power spec-
trum and aperiodic components (Hill et al., 2022; Varela et al., 2001)],
and brain dynamics and interactions [i.e., complexity, and
graph-theoretic measures (Bullmore and Sporns, 2009; Iinuma et al.,
2022)]. We selected these metrics based on their established relevance
to demographic factors, their potential implications for understanding
brain function, and their frequent use across different studies. These
metrics have shown associations with the features assessed in our study,
namely age (Cesnaite et al., 2023; Stacey et al., 2021; Gaal et al., 2010),
sex (Carrier et al., 2001; Pravitha et al., 2005; Miraglia et al., 2015),
education (Wilkinson et al., 2023), and cognition (Finnigan and Rob-
ertson, 2011; Bullmore and Sporns, 2009; Iinuma et al., 2022; McBride
et al., 2014). Power spectrum metrics reveal dominant rhythms linked
with cognitive processes and brain states (Klimesch, 1999), while
spectral EEG metrics interpret non-rhythmic brain activity relevant to
aging and cognition (Donoghue et al., 2020). Complexity metrics pro-
vide insights into the dynamics of different brain regions and their in-
teractions, potentially serving as biomarkers for brain health disorders
(Tononi et al., 1994; Lau et al., 2022). Graph-theoretic metrics illumi-
nate functional connectivity patterns and network organization (Bull-
more and Sporns, 2009). These metrics served as outcomes to evaluate
the predictive value of each demographic and cognitive variable under
two conditions: (a) when it served as the most robust predictor within
the model, and (b) when another variable was considered the best
predictor. Additionally, we explored the source space for each of
brain-phenotype models.
We hypothesized that despite the sample diversity and heterogene-
ity, a subset of the demographic and cognitive variables would signi-
cantly predict each EEG metric. We expected that age (Rosenberg et al.,
2020; Hill et al., 2022; Carrier et al., 2001; Javaid et al., 2022; Merkin
et al., 2023; Meunier et al., 2009; Murty et al., 2020; Smith et al., 2023;
Smits et al., 2016) and cognition (Carrier et al., 2001; Hinault et al.,
2023; Smits et al., 2016; Song et al., 2014) would emerge as the most
robust predictors of EEG metrics. We also anticipated that the combined
effects of these two factors would result in models explaining a high
proportion of variance of EEG measures. The inuence of sex (Merkin
et al., 2023; Stacey et al., 2021; Trammell et al., 2017) and education
(Trondle et al., 2023) would be weaker. Our results suggest that indi-
vidual differences in diverse settings impact core metrics of brain
function that are used in standard case-control studies. The role of in-
dividual differences and diversity in contributing to brain function needs
to be addressed more systematically in global contexts.
2. Materials and methods
2.1. Participants
This multicentric study involved 1298 healthy adult participants
(age: mean =46.60, SD =20.76, range 1891 years;; years of formal
education: mean =13.77, SD =4.38; sex: M =597, F =701) repre-
senting diverse populations (Fig. 1A) from the global south (Argentina,
Brazil, Colombia, Chile, Cuba) and the global north (Ireland, Italy,
Turkey, United Kingdom). Demographic characteristics and sample sizes
for each country are provided in Table 1. Participants had no history of
psychiatric and/or neurological disorders, alcohol/drug abuse, signi-
cant visual and/or auditory impairments. Furthermore, participants
were not using any ferromagnetic implant. No participant reported
subjective cognitive complaints or functional impairments. Data on the
H. Hernandez et al.
NeuroImage 295 (2024) 120636
4
general cognitive state of 725 participants were available. The study
protocol was approved by the Institutional Ethics Committee at each
participating center, and all participants provided written informed
consent in accordance with the Declaration of Helsinki. The data and
analysis codes are freely available at the following GitHub link https
://github.com/euroladbrainlat/Brain-health-in-diverse-setting. The
data in the repository has been anonymized and pre-processed.
We used the G*Power 3.1.9.7 software (Faul et al., 2007) for power
analysis, specically employing an F-test for multiple linear regressions.
Our studys sample size, comprising 725 subjects with cognitive mea-
surements, surpassed the required minimum for conducting multiple
linear regressions with four variables. This allows for the detection of
small effect sizes (0.02 (Cohen, 1988)) with a statistical power of 0.88.
Additionally, the larger total sample of 1298 subjects was adequate for
performing multiple linear regressions with three variables, effectively
identifying small effect sizes with a statistical power of 0.99.
2.2. Demographic and cognitive variables
2.2.1. Demographics
The demographic information includes age at the time of assessment
(in years), sex (male or female), and years of formal education. Data on
gender identity was not available.
2.2.2. Cognition
The general cognitive state was evaluated using the raw total score
from the Mini-Mental State Examination (MMSE) (Folstein et al., 1975).
This test is a widely recognized screening tool for cognitive impairment
and assessing general cognitive functioning. It evaluates various cogni-
tive domains, including orientation to time and place, short-term
memory recall, working memory, language abilities, visuoconstruc-
tional skills, and basic motor commands. This instrument consists of 30
items, and each correct answer adds one point to the total score, which
ranges from 0 to 30. A score of 24 or more indicates normal cognitive
functioning (Creavin et al., 2016; Mukaetova-Ladinska et al., 2022; Cebi
et al., 2020; Foderaro et al., 2022; Kochhann et al., 2010).
Fig. 1. Study design. (A) Sample: Participants were recruited from a multicenter study encompassing regions in the global south (Argentina, Brazil, Chile,
Colombia, Cuba), and the global north (Italy, Ireland, Turkey, United Kingdom). (B) Predictors: The dataset comprised a total of 1298 participants, with cognitive
data available for 725 individuals. Regression models tuned with cross validation and data partition were developed using years of education, age, sex, and cognitive
state as predictors. (C) EEG data preprocessing: The preprocessing steps included electrode re-referencing, noise removal through signal ltering, resampling to 512
Hz, artifact removal (e.g., blinks and eye movements), and transformation into a common source space using the Automated Anatomical Labeling (AAL) atlas. (D)
Outcomes: EEG data in the source space was used to compute four groups of metrics, categorized into power spectrum, aperiodic, complexity, and connectivity
metrics. (E) Data analysis: Linear regressions were applied, utilizing data partitioning with an 80% training sample and a 20% testing set, cross-validated (k =10
repetitions). We reported the importance of predictors and the results on the inuence of each predictor on specic brain regions.
H. Hernandez et al.
NeuroImage 295 (2024) 120636
5
2.3. EEG data acquisition, processing, and harmonization
The EEG acquisition parameters at each center are detailed in Sup-
plementary Table 1. Participants were situated in comfortable chairs
within dimly lit, electromagnetically quiet rooms, and were advised to
stay still and alert. Resting-state EEG (rsEEG) with closed eyes was
recorded using various systems, encompassing distinct sensor types,
calibrations, and electrode congurations (Supplementary Table 1). In
some instances, EEGs with open eyes were also captured. Due to
differing recording durations across centers, EEG analyses were limited
to the rst ve minutes of each recording.
The EEG data was processed ofine using a customized, automated,
and validated pipeline that incorporated harmonization protocols spe-
cically designed to mitigate batch effects and methodological varia-
tions in multi-center EEG studies (Ballesteros et al., 2023; Prado et al.,
2023). The processing workow involved several steps, including data
preprocessing, EEG rescaling, spatial normalizations, and EEG source
localization.
2.3.1. Pre-processing
Raw EEG data was ltered between 0.5 and 40 Hz using an 8th order
zero-phase shift Butterworth lter. Data were re-sampled to 512 Hz and
referenced using the reference electrode standardization technique
(REST) (Hu et al., 2018). Blinking, ocular artifacts and myogenic activity
were corrected using two distinct ICA techniques (Delorme and Makeig,
2004): ICLabel (a tool for the classication of EEG independent com-
ponents into signals and different categories of noise) (Pion-Tonachini
et al., 2019) and EyeCatch (a tool for identifying eye-related ICA scalp
maps) (Bigdely-Shamlo et al., 2013). ICA techniques [Infomax ICA
(G´
orecka and Walerjan, 2011)] were exclusively used to remove ocular
artifacts (blinking and eye movements). These artifacts typically fall
among the rst components of ICA. After ensuring artifact correction
through ICA techniques (specically, ICLabel and EyeCatch), we con-
ducted a visual inspection of the EEG. Malfunctioning channels were
replaced using weighted spherical interpolations (Kothe and Makeig,
2013).
2.3.2. Normalization
The normalization process adhered to multicentric studies guidelines
(Prado et al., 2022). To minimize variability across centers, Z-score
transformations of EEG time series were implemented (Ballesteros et al.,
2023; Prado et al., 2023a, 2023b). The normalization consisted of
computing the mean voltage of each EEG channel and Z-transforming
the corresponding voltage samples, considering the mean and the
standard deviation of the distribution. This normalization reduces the
electrode-by-electrode variability and was conducted independently for
each recruitment center, therefore reducing the inter-site variability.
Furthermore, a spatial normalization of EEG was performed by
combining different electrode congurations, creating virtual electrodes
calculated from topographic interpolation transforms. This method has
been effectively utilized in studies investigating the variance between
acquisition systems in contrast to between-subject and between-session
variances (Melnik et al., 2017). The technique projects electrode posi-
tions onto a mesh-head model consisting of 1082 points and interpolates
EEG activity. We adjusted the EEGLAB headplot (Prado et al., 2023)
function to map the original EEG onto a 6067-point mesh-head model
(Kothe and Makeig, 2013; Melnik et al., 2017).
2.3.3. EEG source estimation
EEG source generators were estimated using the standardized low-
resolution electromagnetic tomography (sLORETA) method (Pascual--
Marqui, 2002). This method estimates the standardized current density
at specic virtual sensors located in the cortical gray matter and the
hippocampus of an average brain (MNI 305, Brain Imaging Centre,
Montreal Neurologic Institute). This estimation is based on a linear,
weighted summation of a unique scalp voltage conguration or the EEG
cross-spectrum at the sensor level. Essentially, sLORETA serves as a
distributed EEG inverse solution technique that extends from a stan-
dardized version of minimum norm current density estimation. It
effectively addresses challenges associated with estimating deep sources
of EEG activity and ensures precise localization, even in cases with
signicant correlation among nearby generators (Asadzadeh et al.,
2020).
Electrode layouts were aligned with the MNI152 scalp coordinates
(Mazziotta et al., 2001). When computing the sLORETA transformation
matrix, a signal-to-noise ratio of 1 was selected as the regularization
method. Standardized current density maps were produced using a three
concentric spheres head model, in a predened source space of 6242
voxels (with a voxel size of 5 ×5 ×5 mm) of the MNI average brain. The
brain was segmented into 82 brain regions using the Automated
Anatomical Labeling (AAL) atlas (Rolls et al., 2015). Standarized current
densities were estimated for each of the 153,600 Vage distributions
comprising the ve-minutes of rsEEG (sampled at 512 Hz). Standarized
current density time series estimated in voxels belonging to the same
AAL regions were averaged, leading to a mean time series for each brain
area (Prado et al., 2023; Cruzat et al., 2023; Herzog et al., 2022).
Table 1
Demographic and cognitive state information.
Country Sex Age Education Cognition
Sex n Mean SD Mean SD n Mean SD
Argentina Female 29 66.76 8.53 17.03 2.03 28 29.04 0.69
Male 12 66.33 7.92 15.42 2.94 12 28.17 1.85
Chile Female 59 57.03 17.10 15.49 3.73 49 28.80 1.47
Male 21 62.62 18.45 14.33 4.35 14 28.30 2.27
Brazil Female 58 59.33 16.88 11.88 3.68 58 28.03 1.61
Male 35 62.69 15.91 13.51 2.34 35 28.29 1.51
Colombia Female 96 49.61 13.85 12.43 4.64 96 27.82 3.03
Male 37 45.49 15.49 12.68 4.59 37 28.00 3.13
Cuba Female 53 37.51 11.34 13.25 2.79 31 29.68 0.70
Male 149 29.68 7366 12.99 2.81 94 29.28 1.42
Italy Female 12 60.55 7.35 15.25 4.85
Male 9 62.90 9.24 15.44 5.46
United Kingdom Female 34 56.82 17.40 15.00 3.67
Male 19 53.89 21.13 15.11 10.27
Ireland Female 41 65.76 4.54 15.50 3.78 40 29.25 0.90
Male 44 69.41 5.85 15.01 4.17 44 28.75 1.10
Turkey Female 319 43.48 22.02 13.37 4.98 118 28.85 1.10
Male 271 39.53 22.03 14.19 4.12 69 29.10 1.36
H. Hernandez et al.
NeuroImage 295 (2024) 120636
6
2.4. EEG descriptors
We computed four categories of EEG descriptors: (i) power spectrum
and (ii) aperiodic spectral metrics, (iii) complexity, and (iv) graph-
theoretic measures. These categories reect neural activity (i.e., power
spectrum or aperiodic spectral metrics) (Cesnaite et al., 2023; Donoghue
et al., 2020), and brain dynamics and interactions (i.e., complexity, and
graph-theoretic measures) (Zappasodi et al., 2015; Vecchio et al., 2022).
Power spectrum metrics are relevant in EEG data analysis, offering in-
sights into brain activity by characterizing how signal power is distrib-
uted across various frequencies (Wang et al., 2015). These metrics
establish a baseline for typical spectral activity, serving as a valuable
benchmark for identifying pathological conditions (Buzs´
aki, 2006) The
aperiodic spectral components of EEG represent non-rhythmic brain
activity. These components facilitate physiological interpretations
related to aging and cognition (Donoghue et al., 2020). For example,
these aperiodic metrics has been linked to excitation/inhibition balance
(Medel et al., 2023; Martinez-Canada et al., 2023), and electrophysio-
logical noise (Voytek et al., 2015). Complexity metrics enable the
evaluation of dynamical brain complexity, which have been explored as
potential biomarkers for diagnosing mental health disorders (Tononi
et al., 1994; Lau et al., 2022). Lastly, graph- theoretic metrics shed light
on the structure of functional connectivity patterns and network orga-
nization (Bullmore and Sporns, 2009).
2.4.1. Power spectrum metrics
Power spectrum analyses were conducted in both the canonical EEG
frequency bands and the subject-specic EEG frequency bands (Babiloni
et al., 2020). Canonical bands provide standardized ranges ensuring
consistency and comparability across studies. However, these bands
may not effectively capture individual differences (Doppelmayr et al.,
1998; Bazanova and Vernon, 2014). Subject-specic bands enhance the
sensitivity to detect specic physiological, cognitive, or behavioral
states that might be masked when using generic bands (Klimesch, 1999).
For instance, the individual alpha frequency peak (IAF) increases during
childhood and slows in middle and older age (Turner et al., 2023), de-
creases with neurodegeneration (Moretti et al., 2004), and increases
with cognitive demands (Haegens et al., 2014). Analyzing power with
canonical and subject-specic frequency bands allows for a compre-
hensive assessment of EEG data.
The canonical frequency bands were dened as follows: delta (δ):
1.56 Hz; theta (θ): 6.58.0 Hz; alpha1 (
α
1): 8.510 Hz; alpha2 (
α
2):
10.512.0 Hz; beta1 (β1): 12.518.0 Hz; beta2 (β2): 18.521.0 Hz; beta3
(β3): 21.530.0 Hz; gamma (γ): 30.040.0 Hz. This EEG band denition
is used for spectral analysis in the EEG source space with LORETA-KEY
software (Grech et al., 2008). It has implemented in many studies for
conducting EEG frequency-specic analyses in dementia (e.g., (Prado
et al., 2023; Aoki et al., 2019)), and other conditions. (e.g. (Lee et al.,
2019; Aoki et al., 2015; Aoki et al., 2019; Ouyang et al., 2020; Krause
et al., 2015)).
Subject-specic frequency bands were determined using the IAF,
dened as the frequency with maximum power in the alpha-frequency
band, and the theta/alpha-frequency transition (TF), dened as the
frequency with minimum power in the second half of theta frequency
range (Martinez-Canada et al., 2023; Babiloni et al., 2020; Ince et al.,
2017). The subject-specic frequency bands are dened as: δ (TF-4 to
TF-2), θ (TF-2 to TF),
α
low
(TF to IAF), and
α
high
(IAF to IAF+2) (Babiloni
et al., 2020). The β and γ frequency bands corresponded to the canonical
division.
We computed both the power spectral density (PSD) and the
normalized PSD (nPSD) using Welchs method with 1-second Hanning
windows with 50% overlap. Subsequently, we calculated the mean of
the nPSD in each frequency band (equivalently percent power) (Li et al.,
2007) and the percent of the nPSD of a given frequency band relative to
the total nPSD (relative power density) (Wang et al., 2015; Babiloni
et al., 2020).
2.4.2. Aperiodic spectral metrics
We computed the aperiodic components of the EEG power spectral
density (PSD): the 1/f slope (a), knee (k), and offset (b) (Martinez-Ca-
nada et al., 2023; Pathania et al., 2021; van Nifterick et al., 2023). We
applied the tting oscillations and one-over-F (FOOOF) algorithm to the
PSD, covering the frequency band from 0.5 to 40 Hz, peak width limits
from 1 to 6 Hz, maximum number of peaks of 6, minimum peak height of
0.2, and peak threshold of 2.0 (Martinez-Canada et al., 2023; Pathania
et al., 2021; van Nifterick et al., 2023). The FOOOF algorithm was
designed to model the aperiodic component of the PSD (Donoghue et al.,
2020), using a Lorentzian function as follows:
A=blog(k+Fx)
where A is the aperiodic component, b is the offset, x is the exponent,
and x=-a.
2.4.3. Complexity metrics
Complexity descriptors of the EEG source space provide insights into
the dynamics of different brain regions and their interactions. A high
complexity value in an EEG signal indicates an increased degree of ir-
regularity, indicating that the EEG is less predictable. This is often
associated with a healthy and alert brain state (Medel et al., 2023).
Conversely, a low comlexity values suggests a more regular or repetitive
signal, could be observed in pathological states (Sun et al., 2020)
anaesthesia or deep sleep (Sarasso et al., 2021; Boncompte et al., 2021).
Fractal dimension (FD), permutation entropy (PE), Wiener Entropy
(WE), and spectral structure variability (SSV) (Sun et al., 2020; Burns
and Rajan, 2015) were computed to assess the complexity of EEG sig-
nals, with code sourced from a GitHub repository (https://github.com/tf
burns/MATLAB-functions-for-complexity-measures-of-one-dimensiona
l-signals) (Burns and Rajan, 2016). We include measures of predict-
ability (FD) as well as regularity (PE, WE, SSV), which represent the two
key aspects of brain dynamics (Lau et al., 2022). Predictability refers to
the ability to anticipate the temporal evolution of the systems states,
while regularity measures the frequency and pattern of repetitions in the
systems trajectory (Lau et al., 2022). Changes in FD are linked to var-
iations (Zappasodi et al., 2015) and cognition (Smits et al., 2016;
Hemmati et al., 2013). Similarly, both PE and WE have been associated
with individual differences in age (Al Zoubi et al., 2018; Shumbaya-
wonda et al., 2018) and cognition (Parbat and Chakraborty, 2021; Seker
et al., 2021). These metrics collectively capture broad conceptions of
complexity (Burns and Rajan, 2015). The equations for complexity
metrics are described below.
Fractional dimension: The FD is a statistic index of complexity
details in a patter and can be expresed as:
FD =a(NLD NLD0)k
where a, k y NLD0, was set to the suggested parameters values: 1.9079,
0.18383 y 0.097178, respectively (Kalauzi et al., 2009). The NLD was
calculate as:
NLD =1
N
N
i=2
|yn(i) yn(i1)|
where yn(i)represents the ith signal sample after amplitude
normalization.
Permutation entropy: The PE is an ordinal-based non-parametric
metric of the temporal dependence structure in linear or non-linear time
series. Therefore, PE can be expressed as follows:
Considering a time series represented by xt, con t=1,,T, and the
embedded vector Xt= [Xt+Xt+l,,+Xt+(n1)], where n is the
embedding dimension and l is the lag. Then, the vector Xt is arranged
from smallest to largest. Finally, PE was dened as:
H. Hernandez et al.
NeuroImage 295 (2024) 120636
7
PE =
n!
n=1
p(
π
)ln(p)
where p(
π
) = f(
π
)/(T (n1)l)y f(
π
)represents the frequency of the
symbols of length n derived from the ordinal relationships between Xt.
(Ouyang et al., 2013)
Wiener entropy and spectral structure variation: The WE, alter-
natively referred to as spectral atness or tonality coefcient, signies
the uniform distribution of signal energy across the frequency domain.
This metric provides a quantiable measure of how closely the signal
mirrors a sinusoidal function, as opposed to exhibiting characteristics
reminiscent of noise. The spectral atness measure (SFM) was used to
quantify spectral structure (Singh, 2011), and the equation for its
calculation is as follows:
SFM(t) = log N
i=1S(t,f)
1
N
1
NN
i=1S(t,f)
where N is the number of points in the Fourier transform, and S(t,f)is
the associated power at each frequency component. Finally, the Wiener
entropy is calculated as the average of SFM(t), and the spectral structure
variability is calculated as the variance of SFM(t).
2.4.4. Connectivity graph metrics
Calculating graph metrics required determining functional connec-
tivity across the entire space of different brain regions. Functional
connectivity was assessed in the time-domain using information theo-
retic measures, which were calculated with custom Matlab codes (Prado
et al., 2023; Herzog et al., 2022). These measures were derived using the
Gaussian copulas approximation (Ince et al., 2017), a robust computa-
tional framework that combines copulas statistical theory with an ana-
lytic solution for the entropy of multivariate Gaussian distributions. The
incorporation of Gaussian copulas-based methods strengthens the
robustness and cost-effectiveness of our methodology, resulting in a
reduction of computational time required for functional connectivity
calculations. We computed both pairwise and high-order interactions
metrics, specically mutual information (MI) (Prado et al., 2023; Herzog
et al., 2022; Santamaria-Garcia et al., 2023), conditional mutual infor-
mation (CMI) (Prado et al., 2023; Ince et al., 2017), and organizational
information (O_info) (Prado et al., 2023) metrics (Supplementary Data
S1). MI gauges the shared information between two random variables
(or time series), while CMI assesses the shared information between two
random variables (or time series) with respect to a specied third vari-
able. Finally, O_info, an extension of Shannons mutual information,
enhances our understanding of the crucial characteristics of multivariate
systems, especially those involving high-order interactions (Herzog
et al., 2022). Information theory metrics surpass traditional measures
like coherence or phase locking value because they can capture
non-linear interactions (Imperatori et al., 2019; Kida et al., 2016) and
provide better results(109). This is crucial for exploring brain dynamics,
where complex non-linear interactions among neural regions shape
patterns of functional connectivity (Imperatori et al., 2019; Kida et al.,
2016).
Measures of segregation, integration, and global metrics were
included to measure complementary dimensions of network topology
(Bullmore and Sporns, 2009; Rubinov and Sporns, 2010). Global Ef-
ciency provides insights into the efciency of information exchange
across the entire network, crucial for global information processing
(Rubinov and Sporns, 2010). Transitivity measures the prevalence of
clustered connectivity, reecting localized, interconnected communities
key for local processing and network segregation (Rubinov and Sporns,
2010). Small-worldness captures the balance between random and
regular networks, indicating an optimal organization of localized and
distributed processing, and integration and segregation, enhancing
network efciency and robustness (Bassett and Bullmore, 2017). Lastly,
density quanties the overall connectivity of the network, offering a
direct assessment of the complexity and connectivity level (Rubinov and
Sporns, 2010). All these measures have been shown to be sensitive to
individual differences in age (Javaid et al., 2022; Meunier et al., 2009;
Achard and Bullmore, 2007; Onoda and Yamaguchi, 2013), sex (Mir-
aglia et al., 2015) and cognition (Finnigan and Robertson, 2011; Bull-
more and Sporns, 2009; Iinuma et al., 2022; McBride et al., 2014; Liao
et al., 2017) . The equations of these metrics are described below.
Connection weight: The connection weight expresses the strength
with which nodes are linked, where wij is the connection weight between
node i and node j.
Weighted shortest path length: The shortest path length is a
fundamental metric to evaluate integration, which is dened as:
dij =
auvgi j
f(wuv)
where f serves as a mapping from weight to length, and gij represents
the shortest weighted path between i and j.
Weighted characteristic path length: The characteristic path
length is a measure employed to evaluate communication efciency in a
weighted network. The equation for its calculation is as follows:
L=1
n
ij=idij
n1
where n represents the number of nodes in the network, and dij is the
length of the shortest path between node i and node j.
Number of triangles: The number of triangles is a fundamental
metric to evaluate segregation, which is dened as:
ti=1
2
j,hwijwih wjh1
3
where ti is the weighted geometric mean of triangles around i, j and h are
indices representing the neighboring nodes of i in the weighted matrix.
Weighted degree: The weighted degree is a measure employed to
evaluate the importance or connectivity of a node in a weighted
network. This metric can be calculated as follows:
ki=
j
wij
where ki is the degree of a node i.
Weighted clustering coefcient: The clustering coefcient is a
metric used to evaluate the tendency of nodes in a network to form local
groupings or clusters. The equation is as follows
C=1
n
i
2ti
ki(ki1)
where ti is the number of triangles around i and ki is the weighted degree
of a node i.
Weighted global efciency: Global efciency was employed to
assess integration, crucial in quantifying the effective sharing of infor-
mation across the brain, essential for coordinated cognitive activity.
This metric evaluates the networks ability to exchange information
globally (Yu et al., 2021; Stanley et al., 2015). The equation for its
calculation is as follows.
E=1
n(n1)
i=j
1
dij
where n represents the number of nodes in the network, and dij is the
length of the shortest path between node i and node j.
Weighted transitivity: Transitivity was selected to measure segre-
gation due to its reection of local interconnectivity, indicating the
H. Hernandez et al.
NeuroImage 295 (2024) 120636
8
tendency of nodes to form closed clusters or triads. It serves as an in-
dicator of local connectivity (Rubinov and Sporns, 2010) and can be
expressed as follow.
T=i2ti
iki(ki1)
where ti is the number of triangles around i and ki is the weighted degree
of a node i.
Weighted density: Density served as global measure, indicating the
overall connectedness of the network addressing complexity and resil-
ience of brain functioning (Rubinov and Sporns, 2010). The equation for
its calculation is as follows.
D=i=jwij
n(n1)
where wij represents the weight of the connection between node i and
node j, and n is the total number of nodes in the network.
Weighted small-worldness: Small-worldness was utilized to eval-
uate both integration and segregation, capturing the optimal balance
between specialized processing within clusters and the global integra-
tion essential for comprehensive brain functionality. A network exhibits
small-world characteristics if it displays a high clustering coefcient
(reecting segregation) and a short characteristic path length (indi-
cating integration) (Rubinov and Sporns, 2010; Bassett and Bullmore,
2017; Yu et al., 2021).
The commonly used measure to quantify small-worldness in a
network includes the clustering coefcient (C) and the characteristic
path length (L). The formula for small-worldness is expressed as the ratio
of the observed clustering coefcient in the network (C) to the expected
clustering coefcient in a random graph with the same number of nodes
and edges (Cr), normalized by the observed characteristic path length in
the network (L) and the expected characteristic path length in the
random graph (Lr).
σ
=C/Cr
L/Lr
A crucial difference between transitivity and the clustering coef-
cient is found in the normalization procedure. Clustering is normalized
individually for each node, while transitivity is normalized across the
entire node set (Rubinov and Sporns, 2010).
2.5. Simplication of the EEG analytical domain
To streamline the analysis of spectral and complexity results from
EEG data, we adopted two complementary approaches. Initially,
anatomically and functionally related AAL regions were merged to form
more consolidated regions of interest (ROIs), as detailed in Supple-
mentary Table 3. Subsequently, we rened the EEG analytical space by
applying specic statistical criteria. The methodology for these pro-
cesses is elaborated in the subsequent sections.
2.5.1. Region selection for analysis
We combined regions from the AAL atlas to generate ten cohesive
ROIs. This approach involved two critical criteria: (i) grouping together
brain regions associated with a specic cortical gyrus (e.g., superior,
middle, and inferior orbital gyri) into a single ROI, and (ii) assembling
neighboring regions with established functional coupling, such as the
Rolandic operculum and insula. The proposed method restructures EEG
analysis by concentrating on ROIs that include both structurally and
functionally related regions, rather than analyzing each small region in
isolation. We utilized a mean averaging approach, calculating the
average value of the metrics derived from the signals of the regions
within the AAL atlas that are associated with each ROI. This simpli-
cation enhances data interpretation and facilitates the identication of
patterns or signicant features in brain activity.
2.5.2. Statistical reduction of spectral and complexity features
To rene the analysis of spectral and complexity metrics, we applied
statistical criteria. Specically, ROIs demonstrating statistically signi-
cant relative power density and equivalent percent power (Wang et al.,
2015; Jeong et al., 2021; Moretti et al., 2004) were identied through
mean-vs-zero non-parametric permutation tests (
α
=0.05; 5000 ran-
domizations (Manly, 1997)). This renement, applied to both canonical
and individual EEG band classications (Babiloni et al., 2020), was
carried out for each frequency band. The results were then adjusted for
multiple comparisons using the Benjamini and Hochberg False Discov-
ery Rate (FDR) method. ROIs representing the Individual Alpha Fre-
quency (IAF) (Klimesch, 1999; Moretti et al., 2004) were those
displaying statistically signicant
α
activities. Similarly, ROIs indicative
of the θ-
α
transition (TF) (Klimesch, 1999; Moretti et al., 2004) were
those with statistically signicant θ activity. Moreover, the analytical
space of aperiodic (Martinez-Canada et al., 2023; Pathania et al., 2021;
van Nifterick et al., 2023) and complexity (Sun et al., 2020; Burns and
Rajan, 2015) metrics was further reduced by employing mean-vs-zero
non-parametric permutation tests, followed by correction using the
Benjamini and Hochberg FDR method.
2.6. Multiple linear regression models
We used multiple linear regression models to understand the rela-
tionship between predictors and outcome variables (Bishop, 2006). The
regression models were constructed using the four predictors outlined in
Section 2.2. We constructed an individual regression model for each
ROI. Regression models were generated and categorized based on the
highest-rated predictors. This approach allowed us to separate the re-
sults into three distinct groups according to the top-rated predictor. The
rst group prioritized age as the top-rated predictor, the second
emphasized education, and the third focused on cognition. None of the
models identied gender as the most important predictor. The mathe-
matical representation of the linear regression model is:
Y=B0+B1x+B1x1+B2x2+Bnxn+ϵ
where Y is the outcome variable, B0 is the intercept, B= [B1,B2,,
Bn]is the coefcients vector of the independent variables matrix X=
[x1,x2,,xn]and ϵ represents the error term. The coefcients B1,B2,
,Bn are determined using the least squares method, which aims to
minimize the sum of the squared differences (errors) between the
observed values (actual values) and the values predicted by the model.
The formulas to calculate the coefcients are:
B=XTX1XTY
where XT is the transpose of matrix of vector X .
The EEG parameters obtained from the chosen ROIs (see Supple-
mentary Table 3) and the connectivity outcomes estimated in whole-
brain analyses were utilized as inputs for the multiple linear regres-
sion models described above.
2.7. Data partition
Models were trained on a training sample (80%) and tested in a
testing set (20%), with k =10 folds (Muller and Guido, 2018). For each
iteration, we computed the estimation coefcients for the predictors,
R-squared, Cohens f
2
, Fishers F of the model, and the models signi-
cance. We reported the mean estimation values for each predictor along
with their standard deviation, average R-squared and Cohens f
2
(Selya
et al., 2012), and overall Fishers F. To determine the overall model
signicance, we combined the models p-values obtained in each iter-
ation using the Fisher method (Fisher, 1992).
H. Hernandez et al.
NeuroImage 295 (2024) 120636
9
2.8. Statistical criteria for the reduction of feature space for frequency
and complexity metrics
A further reduction of the analytical space was conducted using
statistical criteria. To this end, ROIs with statistically signicant relative
power density and equivalent percent power (Wang et al., 2015; Moretti
et al., 2004; Jeong et al., 2021) were selected by implementing
mean-vs-zero non-parametric permutation tests (
α
=0.05; 5000 ran-
domizations (Manly, 1997)). This analysis was conducted for each fre-
quency band, using both canonical and individual (Babiloni et al., 2020)
EEG band classications. Results were corrected for multiple compari-
sons using the Benjamini and Hochberg FDR method (Benjamini and
Hochberg, 1995). Representative ROIs for IAF (Klimesch, 1999; Moretti
et al., 2004) were those for which θ and
α
activities were statistically
signicant. Likewise, ROIs for TF (Klimesch, 1999; Moretti et al., 2004)
were those for which θ activity was statistically signicant. Further-
more, mean-vs-zero non-parametric permutation tests followed by
Benjamini and Hochberg FDR was the method selected for further
reducing the analytical space of the aperiodic (Martinez-Canada et al.,
2023; Pathania et al., 2021; van Nifterick et al., 2023) and complexity
(Sun et al., 2020; Burns and Rajan, 2015) metrics.
2.9. Quality of the signal
We conducted additional analyses to conrm that the number of
electrodes does not impact signal quality and does not interfere with the
observed effects. We calculated the Overall Data Quality (OQD) index in
the source space, using the methodology proposed by Zhao et al. (2023),
and conducted a linear regression to verify that the number of channels
cannot predict signal quality. We used the number of channels as a
predictor and the quality of the nal signals as the outcome. The method
for calculating OQD segmented into 1-second epochs, each labeled as 1
for low-quality epochs or 0 for high-quality epochs. The OQD represents
the percentage of EEG epochs with good quality, ranging from 0 for
signals where all epochs were classied as low quality, to 100 for signals
where all epochs were classied as high quality (Zhao et al., 2023).
3. Results
For each type of EEG descriptor (spectral, complexity, and connec-
tivity EEG outcomes), we categorized the regression models based on
their top-rated predictors, resulting in three distinct groups according to
the best-evaluated predictor. The rst group prioritized age as the best-
evaluated predictor, the second group emphasized education, and the
third group centered on cognition. None of the models identied sex as
the most important predictor. For each group of models and sets of EEG
outcomes, we reported the three models with the highest R
2
values.
Additionally, we reported the adjusted R
2
. In the Supplementary Ma-
terial (Supplementary Figures 13), we provided an assessment of the
model t quality using Q-Q plots and residual vs. tted values plots. In
most cases, the results for the total sample and the subsample with
available cognitive data were similar; therefore, we report the results for
the latter. Further details on the results for the total sample are provided
in Supplementary Tables 1 to 7.
Some participants had portions of the recording with their eyes open
(Supplementary Table 1). Therefore, we analyzed data only from par-
ticipants who underwent eye-close rsEEG exclusively. These results were
comparable to those reported with the entire sample, and are detailed in
Supplementary Tables 8 - 11. In addition, we checked for collinearity
among the predictors (see Supplementary S2 and Supplementary
Figure 4) and conducted supplementary analyses using Ridge re-
gressions (Supplementary Tables 1215), which is a suitable method for
predictors exhibiting collinearity (Tsigler and Bartlett, 2024). The re-
sults obtained with Ridge were comparable to the previous results.
We conducted additional analyses to ensure that variations in the
number of channels across recruitment centers did not inuence the
results. We applied a linear regression to check if the number of channels
predicts signal quality in the source space. The results showed that the
model was not signicant (Supplementary Table 16). Finally, we con-
ducted an analysis of the correlation across the outcomes reported in the
main results (see Supplementary S3, Supplementary Figure 5, and
Supplementary Tables 1720) to assess the effect of their correlation on
the reported effects.
3.1. Age, education and cognition as the best predictors of power spectrum
metrics
Statistical descriptors for all the regression models reported here are
shown in Table 2.
The two regression models predicting age with the highest R
2
values
had subject-specic
α
low
equivalent power as the EEG outcome measure.
The third one had cannonical γ equivalent power. The rst model
reached the highest values in the left occipital region, the second in the
left parietal region, and the third in the left orbitofrontal region
(Table 2A and Fig. 2A). For all three models, the most important and
signicant predictors were age and cognition. In the model for the left
occipital region, sex was the third most important predictor. For the
other two models, education was the third signicant predictor. When
analyzing the total sample, the results for the three models remained
signicant, although R
2
values decreased (Fig. 2A, and Supplementary
Table 4A).
For education as the best rated predictor, the three models with the
highest R
2
values had IAF as the outcome measure. The model with best
scores corresponded with the right medial frontal gyrus, and revealed
education, cognition, age, and sex as signicant predictors (Fig. 2B and
Table 2B). In the second and third models, the highest values were
reached in the left orbitofrontal cortex and the left inferior frontal gyrus,
respectively, education and age were the signicant predictors dis-
playing similar values. Sex and cognition were signicant in the best
model, but not in the other two. In the total sample, R
2
values were low,
but the models remained statistically signicant (Fig. 2B and Supple-
mentary Table 4B).
When cognition was the best-evaluted predictor, the two best models
estimated canon
α
2
relative power metrics more accurately. The third
one performed better in estimating subject-specic
α
low equivalent
power. The rst two models were in the left occipital regions and the
third one in the right parietal region (Fig. 2C and Table 2C). For the
three models, cognition and age emerged as the most important and
signicant predictors. Sex reached signicance in the three models, but
with lower estimates. Education was not statistically signicant in any
model. In the total sample, the models were also statistically signicant,
but showing a decrease in the R
2
values (Fig. 2C and Supplementary
Table 4C).
3.2. Age and cognition as the best predictors of aperiodic spectral metrics
Statistical values for all models reported here are shown in Table 3.
For age as the top predictor, the three best-evaluated models had
slope as the outcome metric. These models were placed in the left
temporal region, the right temporal region, and the left hippocampus,
respectively (Fig. 3A and Table 3A). In all three models, age and
cognition emerged as the most important and signicant predictors. In
the rst model, sex also reached signicance, but not in the second and
the third one. Education was not statistically signicant in any model. In
the total sample, the models were also statistically signicant, but with
lower R
2
values (see Fig. 3A and Supplementary Table 5A).
The best model with cognition as the best-evaluated predictor, esti-
mated slope metrics more accurately. The second and the third models
performed better in estimating offset metrics. The best model of cogni-
tion was in the left occipital region, the second in the left temporal re-
gion and the third in the left hippocampus (Fig. 3C and Table 3B). It is
worth noting that, although reaching signicance, all these models
H. Hernandez et al.
NeuroImage 295 (2024) 120636
10
showed R
2
values lower than 0.1. In the three models, cognition and age
were the most important and signicant predictors. Sex and education
did not reach signicance in any model. In the total sample, the models
were also statistically signicant, but with lower R
2
values (Fig. 3C and
Supplementary Table 5B).
The models in which education was the best-rated predictor were not
statistically signicant (Fig. 3B).
3.3. Cognition as the best predictor of complexity metrics
Statistical values for all models reported here are shown in Table 4.
There were no models where age, education or sex were identied as the
most highly valued predictors.
For cognition as the top predictor, the FD and the WE metrics in the
whole brain were the main outcomes in the two best models. The third
model performed better in spectral structure variation. In all the models,
cognition showed much higher estimates compared to other predictors.
For the best two models, age, and sex were also signicant predictors. In
the third model, sex was signicant. Education did not reach signi-
cance in any model (Table 4A and Fig. 4C). In the total sample, the
models were also statistically signicant but with an important decrease
in R
2
values (Fig. 3C and Supplementary Table 6A).
3.4. Age and education as the best predictors of graph-theoretic measures
Statistical values for all models reported here are shown in Table 5.
When age was the top predictor, the best model estimated transi-
tivity in the CMI matrix more accurately. The second and the third ones
performed better in estimating global efciency and small-worldness in
the MI and CMI matrices, respectively. For all the three models, age and
education were the most important and signicant predictors. In the rst
model, cognition was also signicant. Sex did not reach signicance in
any model (Fig. 4A and Table 5A). In the total sample, the models were
also statistically signicant with almost the same performance as in the
subsample (Fig. 4A and Supplementary Table 7A).
For education as the top predictor, the best model had transitivity as
outcome, the second one global efciency and the third one small-
worldness, all of them in the organizational information. The three
models revealed education and age as the most important and signi-
cant predictors. Cognition and sex were not statistically signicant in
any model (Table 5B and Fig. 4B). The models were also statistically
signicant in the total sample (Fig. 4B and Supplementary Table 7B).
There were no models where cognition was identied as the most
highly valued predictor.
4. Discussion
This study aimed to characterize EEG-derived endophenotypes based
on demographic and cognitive factors across a diverse sample of par-
ticipants. Despite multimodal diversity and heterogeneity (diverse
populations, data acquisition, multicentric assessments, ampliers,
number and type of electrodes), individual differences shaped electro-
physiological brain dynamics. Age emerged as the most robust and
systematic predictor of EEG signals, followed by cognition. Education
and sex were less inuential predictors. Power spectrum activity and
graph-theoretic measures were the most sensitive in capturing individ-
ual differences. Results are relevant for better understanding the indi-
vidual differences that lead to diversity. The use of more affordable and
scalable measures, such as EEG metrics, could be instrumental in
creating future brain charts. The present ndings may challenge tradi-
tional interpretations of case-control differences in brain signatures,
emphasizing the inuence of demographic and cognitive factors in
brain-phenotype associations.
Age was associated with changes power spectrum and aperiodic
spectral activity and less integrated and segregated networks. Age pre-
dicted a decrease in alpha power at parieto-occipital regions (Trondle
et al., 2023; Donoghue et al., 2020), while the opposite direction was
observed for gamma power in orbitofrontal regions (Rempe et al., 2023;
Hunt et al., 2019). This pattern of results could be explained by the
gradual loss of cholinergic function in the basal forebrain with age
Table 2
Results for the power spectrum metrics on the subsample with available
cognitive data.
A. Models with age as the best evaluated feature
Left occipital region (subj spec
α
low
equivalent power), F =129.36, p <1e-15, R
2
=
0.27, R
2
adjusted =0.26, CI =0.07, F
2
=0.36
Predictors Estimate t value p value
Age 0.01592 6.563429 <1e-15
Cognition 0.009309 2.016705 5.81E-08
Sex 0.00224 1.878269 6.25E-07
Education 0.00181 0.470018 0.917864
Left parietal region (subj spec
α
low
equivalent power), F =125.99, p <1e15, R
2
=
0.26, R
2
adjusted =0.26, CI =0.047, F
2
=0.35
Predictors Estimate t value p value
Age 0.01568 6.421747 <1e-15
Cognition 0.008753 1.8797 7.22E-07
Education 0.00388 1.005555 0.074094
Sex 0.00264 2.201016 1.19E-09
Left orbitofrontal region (canonical γ equivalent power), F =120.68, p <1e15, R
2
=
0.25, R
2
adjusted =0.25, CI =0.05, F
2
=0.34
Predictors Estimate t value p value
Age 0.014329 7.140123 <1e-15
Cognition 0.008861 2.333015 8.69E-11
Education 0.00572 1.790197 2.97E-06
Sex 0.000972 0.990132 0.07891
B. Models with education as the best evaluated feature
Right medial frontal gyrus (IAF), F =18.19, p <1e15, R
2
=0.05, R
2
adjusted =0.05,
CI =0.025, F
2
=0.04
Predictors Estimate t value p value
Education 1.01733 1.670948 2.23E-05
Cognition 0.944278 1.283513 0.004605
Age 0.79489 2.075072 1.58E-08
Sex 0.25695 1.36426 0.001602
Left inferior frontal gyrus (IAF), F =12.22, p <1e15, R
2
=0.03, R
2
adjusted =0.03, CI
=0.04, F
2
=0.024
Predictors Estimate t value p value
Education 0.96513 1.353999 0.001883
Age 0.89932 2.00498 6.22E-08
Cognition 0.265095 0.330327 0.990641
Sex 0.19967 0.906448 0.149072
Left orbitofrontal region (IAF), F =9.59, p <1e15, R
2
=0.026, R
2
adjusted =0.02, CI
=0.07, F
2
=0.01
Predictors Estimate t value p value
Education 0.99361 1.475303 0.000392
Age 0.96511 2.275339 2.48E-10
Cognition 0.266543 0.315654 0.996028
Sex 0.15246 0.732682 0.415125
C. Models with cognition as the best evaluated feature
Left occipital region (canonical
α
2
relative power), F =69.25, p <1e15, R
2
=0.16, R
2
adjusted =0.16, CI =0.06, F
2
=0.18
Predictors Estimate t value p value
Cognition 0.000834 2.464738 3.98E-12
Age 0.00071 4.037384 <1e-15
Education 0.0002 0.718862 0.436808
Sex 0.00011 1.299236 0.003882
Left occipital region (canon
α
2
equivalent power), F =69.25, p <1e15, R
2
=0.16, R
2
adjusted =0.16, CI =0.06, F
2
=0.18
Predictors Estimate t value p value
Cognition 0.005837 2.464738 3.98E-12
Age 0.00497 4.037384 <1e-15
Education 0.00141 0.718862 0.436808
Sex 0.00078 1.299236 0.003882
Right occipital region (subject spec
α
high
equivalent power), F =68.55, p <1e15, R
2
=
0.16, R
2
adjusted =0.16, CI =0.09, F
2
=0.19
Predictors Estimate t value p value
Cognition 0.007048 2.334146 9.2E-11
Age 0.00704 4.415218 <1e-15
Sex 0.00134 1.698314 1.38E-05
Education 0.00082 0.347323 0.979016
H. Hernandez et al.
NeuroImage 295 (2024) 120636
11
(Schreckenberger et al., 2004) which is associated with diminished
cholinergic input in the thalamus (Hindriks and van Putten, 2013;
Pfurtscheller and Lopes da Silva, 1999), leading to decreased power in
alpha oscillations (Trondle et al., 2023; Schliebs and Arendt, 2011).
Conversely, increased gamma in the orbitofrontal cortex suggests a
potential compensatory mechanism associated with age-related changes
in frontal lobes, regions susceptible to developmental and aging pro-
cesses (Rempe et al., 2023).
The decrease in the aperiodic slope with increasing age aligns with
previous results (Hill et al., 2022; Merkin et al., 2023; Trondle et al.,
2023) and supports the proposed association between older ages and
increased asynchronous background neuronal ring (Trondle et al.,
2023; Donoghue et al., 2020). As age advances, there is an increase in
random and spontaneous neural activity (Cremer and Zeef, 1987) and a
decrease in the signal-to-noise ratio (Voytek et al., 2015). This reduction
in signal-to-noise may stem from heightened spontaneous/baseline
neural spiking activity (Cremer and Zeef, 1987), disrupting neural
communication delity and potentially contributing to typical
age-related cognitive decline (Voytek et al., 2015). Furthermore, the
ndings of age-related decreases in alpha power and a attened aperi-
odic slope are highly compatible. These two phenomena may reect
aspects of the same neurobiological changes (Trondle et al., 2023).
Decreased thalamic inhibitory control over cortical areas, due to
impaired cholinergic input, is reected in diminished cortical alpha
Fig. 2. Power spectrum metrics results. (A) The three best models, where age was the best-evaluated feature, estimated low-frequency metrics more accurately.
Age and cognition emerged as the most signicant predictors with the highest evaluation. While sex was statistically signicant, its impact in the models was
minimal. When analyzing the total sample, R
2
values decreased, but the predictors remained signicant. (B) The three best models, where education was the best-
evaluated feature, exhibited low R
2
values but maintained statistical signicance. These models IAF metrics more accurately. Both education and age demonstrated
signicance and displayed similar values in the models. Sex and cognition were signicant in the best model. In the total sample, R
2
was very low, but the models
remained statistically signicant. (C) The three best models, where cognition was the best-evaluated feature, had a similar structure to panel A, but they estimated
high and canonical frequencies more accurately. Cognition was the best-evaluated feature, but age reached almost the same value. Education did not reach statistical
signicance. For the total sample, there was a signicant decrease in R
2
. The brains in the right column represent the intensity of predictors in the respective brain
regions for statistically signicant models, for the subsample with cognitive data. The models in occipital regions were the most signicant ones, with parietal regions
also demonstrating signicant models. Some additional regions in the left hemisphere emerged, including the inferior, middle and orbital frontal gyri. The statistical
values for the models with the complete sample are provided within each box.
H. Hernandez et al.
NeuroImage 295 (2024) 120636
12
power, leading to higher cortical excitability, an increased
excitation-to-inhibition ratio, and a higher level of neural noise (Trondle
et al., 2023). These effects, coupled with decreased network integration
(reduced global efciency and small-worldness) (Javaid et al., 2022;
Meunier et al., 2009; Achard and Bullmore, 2007; Onoda and Yama-
guchi, 2013), suggests disruptions in local and network communication
effectiveness. Age was also associated with a loss of functional special-
ization (segregation), manifesting as smaller and more local modules
across brain networks (Song et al., 2014). Thus, aging may be charac-
terized by an increased asynchronous activity and reduced network
integration and segregation.
Better cognition was associated with increased cortical activation,
and decreased asynchronous neural activity. Improved cognition pre-
dicted enhanced information processing capacity, as reected in
increased alpha power across occipital electrodes (Hindriks and van
Putten, 2013; Pfurtscheller and Lopes da Silva, 1999). Better cognition
was also associated with an increase in general spiking activity (offset)
(Pei et al., 2023; Zhang et al., 2023; Waschke et al., 2021) and a decrease
in neural noise (slope) (Smith et al., 2023; Ouyang et al., 2020; Pei et al.,
2023) in occipital, temporal, and hippocampal regions. This aligns with
previous results showing that aperiodic activity reects dynamic ad-
justments of metacognitive states crucial for successful cognitive per-
formance (Zhang et al., 2023). A positive association with whole-brain
entropy and a negative with FD and spectral structure conrmed the
role of complex dynamics in cognition (Iinuma et al., 2022; Parbat and
Chakraborty, 2021). Thus, increased cognitive performance is consis-
tently associated with both local and global dynamics, involving
increased cortical activation, enhanced alpha oscillations, and complex
brain dynamics.
Fewer years of education correlated with higher individual alpha
frequencies in frontal regions and reduced integration and segregation
of brain networks. However, these associations were weaker in com-
parison with other metrics, as reported with fMRI (Raz et al., 2005).
While direct assessments of education as a predictor of EEG metrics have
been limited in prior studies, previous research including education as a
covariate, showed no signicant effects on periodic frequency (da Cruz
et al., 2020) or graph-theoretic measures (Tan et al., 2019). Similarly,
sex did not emerge as the most inuential predictor in any model. Sex
differences have been observed (Carrier et al., 2001; Pravitha et al.,
2005) and not detected in EEG studies (Trondle et al., 2023; Rempe
et al., 2023; Gaubert et al., 2019). Our results suggest that sex is a less
inuential individual predictor, although it inuences signals when
combined with other demographic or cognitive variables. The impact of
education and gender on brain phenotypes should be further investi-
gated by exploring more specic measures capturing population
diversity.
Across all models, periodic frequency and graph-theoretic measures
demonstrated the highest sensitivity in capturing individual differences.
Older age, worse cognition, and being male predicted lower alpha
power, while older age and lower education were associated with less
integrated and segregated networks. Even when combined with age,
education was not found to be an inuential predictor, supporting a
weak relationship between this factor and brain signals (Raz et al., 2005;
da Cruz et al., 2020; Tan et al., 2019). Older age and worse cognition,
when combined, are linked to reduced alpha power and increased neural
noise, aligning with the neural noise hypothesis of aging (Voytek et al.,
2015; Cremer and Zeef, 1987), stating that with increasing age, neural
noise rises, and the reliability of neural communication diminishes,
contributing to cognitive decline. Therefore, age-related cognitive
decline might be attributed to these combined effects, where decreased
alpha oscillatory activity allows for more neural noise, which, in turn,
disrupts neural communication (Voytek et al., 2015). As complexity and
aperiodic components can emerge by intrinsical modulation of brain
dynamics, or as a consequence of metabolic or energetic impairments
(Medel et al., 2023; Kluger et al., 2023), future studies should investi-
gate the inuence of individual differences in demographics and
cognition on changes in these metrics.
It is worth noting that some models demonstrated low R
2
values (R
2
<0.1); the residuals of these models did not show an approximately
normal distribution and exhibited poor t. However, due to the intrinsic
complexity of neuroimaging data and the inherent variability in brain-
behavior associations, even models with low R
2
can provide valuable
insights into the effects among variables. Additionally, the variation in
R
2
values provide important information about the strongest effects.
4.1. Limitations
Several limitations of our study should be acknowledged. Cognitive
data were only available for a subsample of participants. Additionally,
cognition was assessed using a screening tool. Although the MMSE has
been widely used as a reliable measure of general cognitive state (Fol-
stein et al., 1975), it may not fully encompass the spectrum of cognitive
abilities. Contrary to prior ndings associating graph metrics with
cognition (Bullmore and Sporns, 2009; Yu et al., 2021; Stanley et al.,
2015), our results indicated that MMSE scores did not possess strong
predictive value. Limited performance variability in MMSE scores, with
all participants scoring above 24, affects the range and reduces vari-
ability. We acknowledge the limitations of the MMSE as a tool for
assessing cognition, particularly in healthy populations. However, we
Table 3
Results for aperiodic spectral metrics on the subsample with available cognitive
data.
A. Models with age as the best evaluated feature
Left temporal region (slope), F =50.12, p <1e15, R
2
=0.12, R
2
adjusted =0.12, CI =
0.06, F
2
=0.13
Predictors Estimate t value p value
Age 1.43333 3.930518 <1e-15
Cognition 0.856312 1.208334 0.01079
Education 0.24692 0.469671 0.900581
Sex 0.23427 1.30627 0.003451
Right temporal region (slope), F =41.16, p <1e15, R
2
=0.10, R
2
adjusted =0.10, CI
=0.06, F
2
=0.11
Predictors Estimate t value p value
Age 1.30515 3.67437 <1e-15
Cognition 0.877152 1.278449 0.005588
Education 0.18117 0.338246 0.986777
Sex 0.16609 0.947864 0.117285
Left hippocampus (slope), F =37.40, p <1e15, R
2
=0.10, R
2
adjusted =0.09, CI =
0.05, F
2
=0.10
Predictors Estimate t value p value
Age 1.27768 3.474221 <1e-15
Cognition 0.896495 1.265951 0.006104
Education 0.136882 0.272953 0.999229
Sex 0.13321 0.732632 0.425399
B. Models with cognition as the best evaluated feature
Left occipital (slope), F =26.55, p <1e15, R
2
=0.07, R
2
adjusted =0.07, CI =0.025,
F
2
=0.07
Predictors Estimate t value p value
Cognition 1.226913 1.675558 2.23E-05
Age 1.01193 2.689972 1.39E-14
Education 0.11348 0.282464 0.998198
Sex 0.04598 0.324079 0.991951
Left Temporal (offset), F =22.15, p <1e15, R
2
=0.06, R
2
adjusted =0.06, CI =0.03,
F
2
=0.05
Predictors Estimate t value p value
Cognition 1.931183 1.398192 0.001154
Age 1.76673 2.492283 1.74E-12
Education 0.61812 0.574947 0.792155
Sex 0.35667 1.023012 0.061897
Left hippocampus (offset), F =18.55, p <1e15, R
2
=0.05, R
2
adjusted =0.05, CI =
0.03, F
2
=0.05
Predictors Estimate t value p value
Cognition 2.300525 1.679279 2.35E-05
Age 1.58893 2.239911 5.16E-10
Education 0.167667 0.158416 0.999986
Sex 0.12838 0.36527 0.981173
H. Hernandez et al.
NeuroImage 295 (2024) 120636
13
employed the MMSE because it remains a valuable cognitive screening
in both clinical and research settings. Its widespread availability, ease of
use, and brevity render it a useful tool for initial cognitive assessment in
environments necessitating rapid and accessible evaluations (Creavin
et al., 2016). Further assessments should explore the role of individual
differences in cognition using more comprehensive measures across
various cognitive domains. Moreover, the measure of years of education
may not be sufciently sensitive to capture population diversity in this
factor. Future studies should consider incorporating more detailed
measures that provide information on the quality and nature of educa-
tion received by participants. As we did not inquire about gender
identities, a factor that can signicantly inuence diversity among
populations, future studies should systematically address individual
differences related to gender identities to provide a more comprehensive
understanding of the associated brain signatures. Also, although
country-level analyses are highly relevant, we included only a limited
number of nations with unbalanced sample sizes, thereby reducing the
possibilities for cross-country interpretations. Although these effects are
beyond the scope of this work, future global approaches with larger,
balanced, and more diverse samples should explore country-level effects
effectively.
It is also worth noting that the choice of parceling scheme and
methodology for dening the ROIs can inuence the outcomes. In our
study, we opted to employ ROIs for two primary reasons: (i) To mitigate
potential effects resulting from employing different electrode congu-
rations and quantities, given the EEGs low spatial resolution and the
Fig. 3. Results on the models using aperiodic spectral metrics as outcomes. (A) The three best models when age was the best-evaluated feature identied age
and cognition as the metrics with the highest evaluation, both of which were statistically signicant. Education and sex did not reach statistical signicance. In the
total sample, R
2
experienced a decrease, but the models remained statistically signicant. (B) The models when education was the best-evaluated feature were
presented with transparency because none of them were statistically signicant. (C) The three best models when cognition was the best-evaluated feature had an R
2
lower than 0.1, but they were still statistically signicant. Education and Sex did not attain statistical signicance in any model. In the total sample, the models
experienced a signicant decrease in R
2
. The brains in the right column represent the intensity of predictors in the respective brain regions for statistically signicant
models, for the subsample with cognitive data. The models in hippocampus and temporal regions were the most signicant ones. The statistical values for the models
with the complete sample are provided within each box.
H. Hernandez et al.
NeuroImage 295 (2024) 120636
14
variance in estimation error that arises from using low and high-density
electrode congurations. This approach facilitates comparison of results
across a broad spectrum of electrode arrangements and quantities. (ii)
To reduce the number of regressions. Despite this reduction, we still had
to construct a substantial number of models. On the other hand, using
ROIs requires nding consistent effects in broad regions, and as our
results demonstrate, they displayed robust effects. Future studies using
only high-density arrays should explore the impact of spatial variation
using ner areas. In adittion, we observed variations in the number of
channels across different centers, ranging from 21 to 132 electrodes. To
address the potential effects of these variations, we employed a mesh
model approach for integrating the electrode layouts (Melnik et al.,
2017). We applied spatial normalization to ensure that the EEG inver-
sion process remains unaffected by the number of channels (Prado et al.,
2023). This was veried by conrming that the number of channels did
not predict the signal quality in the source space. Furthermore, our re-
sults were reported with an MNI average brain for source estimation,
given that individual MRI data was unavailable. Source space provides
more adequate spatial resolution with 64+channels. Although chal-
lenging, source analysis can be efciently done with low-density elec-
trodes (Nguyen-Danse et al., 2021; Soler et al., 2020; Baroumand et al.,
2018). We constructed a regression model to predict signal quality based
on the number of channels, which yielded non-signicant results, sug-
gesting that the number of channels did not impact the reported asso-
ciations. We avoided analyzing small regions and focus only on larger
effects. However, using the MNI average brain for source estimation
with low electrode density represents a limitation. Future studies should
conduct further analyses of the impact of the number of channels on
source variability and its associations with brain phenotypes.
Moreover, the correlation among the reported outcomes exhibited
expected values. Metrics that were spatially close showed high levels of
correlation, as did the connectivity metrics. However, the regressions of
predictors with the outcomes were independent of the relationships
among the outcomes. We opt to use different measures as is typical in the
literature (Cesnaite et al., 2023; Zappasodi et al., 2015; Donoghue et al.,
2020; Vecchio et al., 2022) and emphasize their individual effects.
Finally, we identied collinearity between cognition and education in
the sample with cognitive data, while there were no collinearity in the
complete sample. Collinearity can cause issues with estimated accuracy,
high coefcient variances, interpretation challenges, among others
(Snee, 1983). However, our models showed a high level of consistency
and precision in their estimates across both samples (sample with
cognitive data and the total sample). The results yielded the same effects
when employing a collinearity-robust model such as Ridge regression
(Tsigler and Bartlett, 2024).
4.2. Implications and future directions
Despite the diversity of our sample and the heterogeneity in data
acquisition across centers, our results revealed that demographic and
cognitive factors robustly predicted EEG modulations. These ndings
reect the potential of using more affordable and scalable techniques,
such as EEG, for the study of individual differences contributing to di-
versity and brain signatures. Although we did not include EEG micro-
states in this work, as it was beyond the scope of our study, previous
reports have shown sex-specic changes in microstate dynamics during
adolescence as well as at older age (Tomescu et al., 2018). Additionally,
changes in microstates have been associated with age, from childhood
(Hill et al., 2023) to adulthood (Koenig et al., 2002). As EEG microstates
can inform the temporal dynamics of large-scale brain networks across
millisecond timescales, future studies should use them to explore the
predictive value of demographic and cognitive factors across large and
diverse populations. Similarly, given that other complexity metrics like
multiscale entropy have shown sensitivity to variations in age (Wang
et al., 2016) and cognition (Maturana-Candelas et al., 2019), future
research should leverage these tools in large and diverse samples.
Future studies should employ EEG metrics in normative modeling to
create brain charts as anchor points for standardized quantication of
brain functioning over the lifespan, considering individual differences
contributing to population diversity. Normative modeling with neuro-
imaging measures (Bethlehem et al., 2022; Elad et al., 2021; Di Biase
et al., 2023; Rutherford et al., 2022; Habes et al., 2021; Rosenberg et al.,
2020) has claried individual differences in the context of brain devel-
opment or aging, brain health and disease, and mapping variations
across multiple cognitive domains. The development of has brain charts
challenged traditional interpretations of case-control differences (Mar-
quand et al., 2016), which become problematic in domains such as
neurodegeneration and psychiatry where disorders are diagnosed based
on symptoms that overlap between disorders, often yielding heteroge-
neous clinical groups. As normative modeling does not require cate-
gorical partitioning, applying this approach using EEG signals will allow
for a more global applicability in parsing heterogeneity and diversity in
brain health, psychiatry, and neurodegeneration.
Our ndings endorse EEG as a potential tool that, in the future,
should be critically incorporated into case-control studies in clinical
settings (Rossini et al., 2020). While EEG offers numerous benets for
investigating physiological changes linked to pathology, it often relies
on group-level data to identify associations or differences, which limits
its utility at the individual diagnostic level. Established methods like
MRI, PET, and cerebral spinal uid analysis offer undeniable value in
individual diagnosis; however, they are often expensive and have
limited accessibility, particularly in resource-constrained settings. EEG
offers complementary accessible tools that can aid in revealing physio-
logical changes linked to early pathology and the risk of dementia
(Rossini et al., 2020; Parra, 2022).
4.3. Conclusion
Variations in EEG metric typically used in case-control studies are
inuenced by individual differences in demographics and cognition.
Older age, poorer cognition, and being male were associated with
reduced alpha power, whereas older age and less education were linked
to less integrated and segregated networks. Moreover, older age and
worse cognition were linked to reduced alpha power and increased
neural noise. Our ndings pave the way for the future use of EEG-based
brain charts for standardized quantication of brain functioning over
the lifespan, considering individual differences contributing to
Table 4
Results for complexity metrics on the subsample with available cognitive data.
A. Models with cognition as the best evaluated feature
Whole brain (FD), F =55.40, p <1e15, R
2
=0.13, R
2
adjusted =0.13, CI =0.08, F
2
=
0.16
Predictors Estimate t value p value
Cognition 0.42149 2.832611 <1e-15
Age 0.177473 2.26391 3.4E-10
Sex 0.094606 2.451106 4.67E-12
Education 0.06247 0.497431 0.862861
Whole brain (WE), F =39.14, p <1e15, R
2
=0.10, R
2
adjusted =0.10, CI =0.042, F
2
=0.11
Predictors Estimate t value p value
Cognition 0.468037 3.762755 <1e-15
Sex 0.07337 2.284629 1.97E-10
Age 0.072522 1.117352 0.02743
Education 0.018191 0.226368 0.999692
Left inferior frontal gyrus (spectral structure variation), F =36.81, p <1e15, R
2
=
0.09, R
2
adjusted =0.09, CI =0.04, F
2
=0.10
Predictors Estimate t value p value
Cognition 0.15205 3.071085 <1e-15
Sex 0.030633 2.395504 1.67E-11
Education 0.00804 0.24164 0.999116
Age 9.56E-05 0.225578 0.999868
H. Hernandez et al.
NeuroImage 295 (2024) 120636
15
Fig. 4. Results on the models using connectivity and complexity metrics as outcomes. In terms of connectivity metrics, cognition never emerged as the best-
evaluated feature, whereas in complexity metrics, neither age nor education stood out as the top-rated predictors. (A) The three best models when age was the
primary feature exhibited signicant results with high R
2
values and effect sizes for measures of segregation, integration, and small-worldness. Age and education
were the best evaluated and signicant predictors in the models. In the total sample, the models maintained almost the same performance. (B) The three best models
when education was the primary feature yielded a low R
2
but remained statistically signicant. Education and age were the best-evaluated predictors, whereas
cognition and sex did not hold signicance. (C) Results for complexity metrics. The three best models when cognition was the primary feature for predicting
connectivity metrics emerged with several complexity metrics: fractal dimension, spectral structure variability, and Wiener entropy. Cognition had a much higher
value compared to other predictors. In the total sample, the models signicantly decreased R
2
values (see values inside the box). Brains in this panel indicate that the
most predominant regions were the whole brain, along with the occipital, inferior frontal gyrus, and temporal lobe. These regions prevailed in both hemispheres. The
brains to the right of panels (A) and (B) represent the utilized connectivity metrics for the subsample with cognitive data. The statistical values for the models with
the complete sample are provided within each box.
H. Hernandez et al.
NeuroImage 295 (2024) 120636
16
population diversity. Such an approach will allow for more tailored
global models to understanding the variations and diversity in brain
health, psychiatry, and neurology.
CRediT authorship contribution statement
Hernan Hernandez: Writing review & editing, Writing original
draft, Methodology, Investigation, Formal analysis, Conceptualization.
Sandra Baez: Writing review & editing, Writing original draft,
Investigation, Formal analysis, Conceptualization. Vicente Medel:
Writing review & editing, Supervision, Investigation, Conceptualiza-
tion. Sebastian Moguilner: Writing review & editing, Methodology,
Investigation, Formal analysis, Conceptualization. Jhosmary Cuadros:
Writing review & editing, Supervision, Investigation, Data curation.
Hernando Santamaria-Garcia: Writing review & editing, Supervi-
sion, Investigation, Conceptualization. Enzo Tagliazucchi: Writing
review & editing, Supervision, Investigation, Conceptualization. Pedro
A. Valdes-Sosa: Writing review & editing, Supervision, Investigation,
Conceptualization. Francisco Lopera: Writing review & editing, Su-
pervision, Investigation, Conceptualization. John Fredy OchoaG´
omez:
Writing review & editing, Supervision, Investigation. Alfredis
Gonz´
alez-Hern´
andez: Writing review & editing, Supervision, Inves-
tigation. Jasmin Bonilla-Santos: Writing review & editing,
Supervision, Investigation. Rodrigo A. Gonzalez-Montealegre:
Writing review & editing, Supervision, Investigation. Tuba Aktürk:
Writing review & editing, Supervision, Investigation. Ebru Yıldırım:
Writing review & editing, Supervision, Investigation. Renato Anghi-
nah: Writing review & editing, Supervision, Investigation, Data
curation. Agustina Legaz: Writing review & editing, Supervision,
Investigation, Data curation. Sol Fittipaldi: Writing review & editing,
Supervision, Investigation. G¨
orsev G. Yener: Writing review & edit-
ing, Supervision, Resources, Investigation, Funding acquisition. Javier
Escudero: Writing review & editing, Supervision, Investigation.
Claudio Babiloni: Writing review & editing, Supervision, Investiga-
tion. Susanna Lopez: Writing review & editing, Supervision, Investi-
gation. Robert Whelan: Writing review & editing, Supervision,
Investigation, Conceptualization. Alberto A Fern´
andez Lucas: Writing
review & editing, Supervision, Investigation. Adolfo M. García:
Writing review & editing, Supervision, Investigation. David Huepe:
Writing review & editing, Supervision, Investigation. Gaetano Di
Caterina: Writing review & editing, Supervision, Investigation. Mar-
cio Soto-A˜
nari: Writing review & editing, Validation, Investigation.
Agustina Birba: Writing review & editing, Supervision, Investigation.
Agustin Sainz-Ballesteros: Writing review & editing, Supervision,
Investigation. Carlos Coronel: Writing review & editing, Validation,
Investigation. Eduar Herrera: Writing review & editing, Supervision,
Investigation. Daniel Abasolo: Writing review & editing, Supervision,
Investigation. Kerry Kilborn: Writing review & editing, Supervision,
Investigation. Nicol´
as Rubido: Writing review & editing, Supervision,
Investigation. Ruaridh Clark: Writing review & editing, Supervision,
Investigation. Ruben Herzog: Writing review & editing, Supervision,
Investigation. Deniz Yerlikaya: Writing review & editing, Supervi-
sion, Investigation. Bahar Güntekin: Writing review & editing, Su-
pervision, Investigation. Mario A. Parra: Writing review & editing,
Supervision, Resources, Investigation, Funding acquisition. Pavel
Prado: Writing review & editing, Supervision, Investigation. Agustin
Ibanez: Writing review & editing, Writing original draft, Supervi-
sion, Methodology, Investigation, Funding acquisition,
Conceptualization.
Declaration of competing interest
none.
Data availability
Both the data and analysis codes are openly accessible through the
following GitHub link https://github.com/euroladbrainlat/Brain-
health-in-diverse-setting.
Acknowledgments
This work was supported by the Latin American Brain Health Insti-
tute (BrainLat) Seed Grant BL-SRGP2020-02 awarded to MAP and AI.
AMG is an Atlantic Fellow at the Global Brain Health Institute (GBHI)
and is partially supported with funding from the National Institute On
Aging of the National Institutes of Health (R01AG075775,
2P01AG019724); ANID (FONDECYT Regular 1210176, 1210195);
GBHI, Alzheimers Association, and Alzheimers Society (Alzheimers
Association GBHI ALZ UK-22-865742); the Latin American Brain Health
Institute (BrainLat), Universidad Adolfo Ib´
a˜
nez, Santiago, Chile (#BL-
SRGP2021-01); Programa Interdisciplinario de Investigaci´
on Experi-
mental en Comunicaci´
on y Cognici´
on (PIIECC), Facultad de Human-
idades, USACH. AI is supported by grants from the MULTI-PARTNER
CONSORTIUM TO EXPAND DEMENTIA RESEARCH IN LATIN AMER-
ICA [ReDLat, supported by Fogarty International Center (FIC), National
Institutes of Health, National Institutes of Aging (R01 AG057234, R01
AG075775, R01 AG21051, R01 AG083799, CARDS-NIH), Alzheimers
Table 5
Results for graph- theoretic measures on the subsample with available cognitive
data.
A. Models with age as the best evaluated feature
Transitivity (CMI), F =90.62, p <1e15, R
2
=0.20, R
2
adjusted =0.20, CI =0.062, F
2
=0.26]
Predictors Estimate t value p value
Age 0.16318 5.923662 <1e-15
Education 0.06215 1.44298 0.000756
Cognition 0.059867 1.105595 0.032926
Sex 0.0124 0.921539 0.138854
Global efciency (MI), F =88.94, p <1e15, R
2
=0.20, R
2
adjusted =0.20, CI =0.045,
F
2
=0.24
Predictors Estimate t value p value
Age 0.197 6.135292 <1e-15
Education 0.0807 1.563886 0.000111
Cognition 0.037216 0.576938 0.756623
Sex 0.00805 0.509363 0.870492
Small worldness (CMI), F =81.90, p <1e15, R
2
=0.19, R
2
adjusted =0.18, CI =0.06,
F
2
=0.23
Predictors Estimate t value p value
Age 0.14729 5.520251 <1e-15
Education 0.05611 1.346702 0.00252
Cognition 0.050124 0.952187 0.115618
Sex 0.01036 0.791281 0.315885
B. Models with education as the best evaluated feature
Transitivity (O_info), F =8.40, p =0.026, R
2
=0.02, R
2
adjusted =0.02, CI =0.04, F
2
=0.01
Predictors Estimate t value p value
Education 0.002819 1.238426 0.008346
Age 0.00281 1.977057 1.33E-07
Cognition 0.00181 0.645702 0.717019
Sex 0.000173 0.257998 0.999564
Global efciency (O_info), F =8.04, p =0.014, R
2
=0.03, R
2
adjusted =0.02, CI =
0.04, F
2
=0.024
Predictors Estimate t value p value
Education 0.001219 1.355786 0.00202
Age 0.001002 1.79685 3.16E-06
Cognition 0.00065 0.571754 0.848408
Sex 4.97E-05 0.246322 0.99862
Small worldness (O_info), F =7.30, p =0.017, R
2
=0.02, R
2
adjusted =0.02, CI =
0.04, F
2
=0.02
Predictors Estimate t value p value
Education 0.000295 1.412363 0.001037
Age 0.00025 1.921802 3.78E-07
Cognition 0.0002 0.758745 0.468241
Sex 1.49E-05 0.267985 0.998259
H. Hernandez et al.
NeuroImage 295 (2024) 120636
17
Association (SG-20-725707), Rainwater Charitable Foundation The
Blueeld project to cure FTD, and Global Brain Health Institute)], USS-
FIN-23-FAPE-09, ANID/FONDECYT Regular (1210195 and 1210176
and 1220995); ANID/FONDAP/15150012; ANID/PIA/ANILLOS
ACT210096; FONDEF ID20I10152, and ANID/FONDAP 15150012. The
contents of this publication are solely the responsibility of the authors
and do not represent the ofcial views of these institutions.
Supplementary materials
Supplementary material associated with this article can be found, in
the online version, at doi:10.1016/j.neuroimage.2024.120636.
References
Achard, S., Bullmore, E., 2007. Efciency and cost of economical brain functional
networks. PLoS. Comput. Biol. 3 (2), e17.
Alladi, S., Hachinski, V., 2018. World dementia: one approach does not t all. Neurology.
91 (6), 264270.
Allouh, M.Z., Al Barbarawi, M.M., Ali, H.A., Mustafa, A.G., Alomari, S.O, 2020.
Morphometric analysis of the corpus callosum according to age and sex in middle
eastern arabs: racial comparisons and clinical correlations to Autism Spectrum
Disorder. Front. Syst. Neurosci. 14, 30.
Al Zoubi, O., Ki Wong, C., Kuplicki, R.T., Yeh, H.W., Mayeli, A., Refai, H., et al., 2018.
Predicting Age From Brain EEG Signals-A. Mach. Learn. Approach. Front. Aging
Neurosci. 10, 184.
Aoki, Y., Ishii, R., Pascual-Marqui, R.D., Canuet, L., Ikeda, S., Hata, M., et al., 2015.
Detection of EEG-resting state independent networks by eLORETA-ICA method.
Front. Hum. Neurosci. 9, 31.
Aoki, Y., Kazui, H., Pascal-Marqui, R.D., Ishii, R., Yoshiyama, K., Kanemoto, H., et al.,
2019a. EEG Resting-State Networks in Dementia with Lewy Bodies Associated with
Clinical Symptoms. Neuropsychobiology. 77 (4), 206218.
Aoki, Y., Kazui, H., Pascual-Marqui, R.D., Ishii, R., Yoshiyama, K., Kanemoto, H., et al.,
2019b. EEG resting-state networks responsible for gait disturbance features in
idiopathic normal pressure hydrocephalus. Clin. EEG. Neurosci. 50 (3), 210218.
Aranda, M.P., Kremer, I.N., Hinton, L., Zissimopoulos, J., Whitmer, R.A., Hummel, C.H.,
et al., 2021. Impact of dementia: health disparities, population trends, care
interventions, and economic costs. J. Am. Geriatr. Soc. 69 (7), 17741783.
Asadzadeh, S., Youse, Rezaii T, Beheshti, S., Delpak, A., Meshgini, S, 2020. A systematic
review of EEG source localization techniques and their applications on diagnosis of
brain abnormalities. J. Neurosci. Methods 339, 108740.
Babiloni, C., Barry, R.J., Basar, E., Blinowska, K.J., Cichocki, A., Drinkenburg, W., et al.,
2020. International federation of clinical neurophysiology (IFCN) - EEG research
workgroup: recommendations on frequency and topographic analysis of resting state
EEG rhythms. Part 1: applications in clinical research studies. Clin. Neurophysiol.
131 (1), 285307.
Baez, S., Alladi, S., Ibanez, A., 2023. Global South research is critical for understanding
brain health, ageing and dementia. Clin. Transl. Med. 13 (11).
Ballesteros, A.S., Prado, P., Ibanez, A., Perez, J.A.M., Moguilner, S., 2023. A Pipeline for
Large-scale Assessments of Dementia EEG Connectivity Across Multicentric Settings.
In: Preprints, O (Ed.), A Pipeline for Large-scale Assessments of Dementia EEG
Connectivity Across Multicentric Settings. editor.
Baroumand, A.G., van Mierlo, P., Strobbe, G., Pinborg, L.H., Fabricius, M., Rubboli, G.,
et al., 2018. Automated EEG source imaging: a retrospective, blinded clinical
validation study. Clin. Neurophysiol. 129 (11), 24032410.
Bassett, D.S., Bullmore, E.T., 2017. Small-World Brain Networks Revisited.
Neuroscientist. 23 (5), 499516.
Bazanova, O.M., Vernon, D., 2014. Interpreting EEG alpha activity. Neurosci. Biobehav.
Rev. 44, 94110.
Benjamini, Y., Hochberg, Y., 1995. Controlling the false discovery rate: a practical and
powerful approach to multiple testing. J. R. Stat. Soc.: Series B (Methodological) 57
(1), 289300.
Bethlehem, R.A.I., Seidlitz, J., White, S.R., Vogel, J.W., Anderson, K.M., Adamson, C.,
et al., 2022. Brain charts for the human lifespan. Nature 604 (7906), 525533.
Bigdely-Shamlo, N., Kreutz-Delgado, K., Kothe, C., Makeig, S., 2013. EyeCatch: data-
mining over half a million EEG independent components to construct a fully-
automated eye-component detector. Annu Int. Conf. IEEe Eng. Med. Biol. Soc. 2013,
58455848.
Bigdely-Shamlo, N., Mullen, T., Kothe, C., Su, K.M., Robbins, K.A., 2015. The PREP
pipeline: standardized preprocessing for large-scale EEG analysis. Front.
Neuroinform. 9, 16.
Bishop, C.M., 2006. Pattern Recognition and Machine Learning. Springer, New York
[2006]©2006.
Boncompte, G., Medel, V., Cortínez, L.I., Ossand´
on, T., 2021. Brain activity complexity
has a nonlinear relation to the level of propofol sedation. Br. J. Anaesth. 127 (2),
254263.
Bullmore, E., Sporns, O., 2009. Complex brain networks: graph theoretical analysis of
structural and functional systems. Nat. Rev. Neurosci. 10 (3), 186198.
Burns T., Rajan R. A mathematical approach to correlating objective spectro-temporal
features of environmental sounds with their subjective perceptions2016.
Burns, T., Rajan, R., 2015. Combining complexity measures of EEG data: multiplying
measures reveal previously hidden information. F1000Res. 4, 137.
Buzs´
aki, G., 2006. Rhythms of the Brain. Oxford University Press. https://doi.org/
10.1093/acprof:oso/9780195301069.001.0001. Available from.
Carrier, J., Land, S., Buysse, D.J., Kupfer, D.J., Monk, T.H., 2001. The effects of age and
gender on sleep EEG power spectral density in the middle years of life (ages 20-60
years old). Psychophysiology. 38 (2), 232242.
Cebi, M., Babacan, G., Oktem, Tanor O, Gurvit, H, 2020. Discrimination ability of the
Short Test of Mental Status (STMS) compared to the Mini Mental State Examination
(MMSE) in the spectrum of normal cognition, mild cognitive impairment, and
probable Alzheimers disease dementia: the Turkish standardization study. J. Clin.
Exp. Neuropsychol. 42 (5), 450458.
Cesnaite, E., Steinfath, P., Jamshidi, Idaji M, Stephani, T., Kumral, D., Haufe, S., et al.,
2023. Alterations in rhythmic and non-rhythmic resting-state EEG activity and their
link to cognition in older age. Neuroimage 268, 119810.
Cohen J. Statistical Power Analysis for the Behavioral Sciences: L. Erlbaum Associates;
1988.
Cole, J.H., Ritchie, S.J., Bastin, M.E., Valdes, Hernandez MC, Munoz Maniega, S.,
Royle, N., et al., 2018. Brain age predicts mortality. Mol. Psychiatry 23 (5),
13851392.
Cover T.M., Thomas J.A. Entropy, relative entropy and mutual information. Elements
Inform. Theory. 2nd edition 2005. p. 1355.
Creavin, S.T., Wisniewski, S., Noel-Storr, A.H., Trevelyan, C.M., Hampton, T.,
Rayment, D., et al., 2016. Mini-Mental State Examination (MMSE) for the detection
of dementia in clinically unevaluated people aged 65 and over in community and
primary care populations. Cochrane Database Syst. Rev. 2016 (1), CD011145.
Cremer, R., Zeef, E.J., 1987. What kind of noise increases with age? J. Gerontol. 42 (5),
515518.
Cruzat, J., Herzog, R., Prado, P., Sanz-Perl, Y., Gonzalez-Gomez, R., Moguilner, S., et al.,
2023. Temporal Irreversibility of Large-Scale Brain Dynamics in Alzheimers
Disease. J. Neurosci. 43 (9), 16431656.
da Cruz, J.R., Favrod, O., Roinishvili, M., Chkonia, E., Brand, A., Mohr, C., et al., 2020.
EEG microstates are a candidate endophenotype for schizophrenia. Nat. Commun.
11 (1), 3089.
Delorme, A., Makeig, S., 2004. EEGLAB: an open source toolbox for analysis of single-
trial EEG dynamics including independent component analysis. J. Neurosci. Methods
134 (1), 921.
Di Biase, M.A., Tian, Y.E., Bethlehem, R.A.I., Seidlitz, J., Alexander-Bloch, A.F., Yeo, B.T.
T., et al., 2023. Mapping human brain charts cross-sectionally and longitudinally.
Proc. Natl. Acad. Sci. U. S. A. 120 (20), e2216798120.
Dimitriadis, S.I., 2022. Universal Lifespan Trajectories of Source-Space Information Flow
Extracted from Resting-State MEG Data. Brain Sci. 12 (10).
Donoghue, T., Haller, M., Peterson, E.J., Varma, P., Sebastian, P., Gao, R., et al., 2020b.
Parameterizing neural power spectra into periodic and aperiodic components. Nat.
Neurosci. 23 (12), 16551665.
Donoghue, T., Haller, M., Peterson, E.J., Varma, P., Sebastian, P., Gao, R., et al., 2020a.
Parameterizing neural power spectra into periodic and aperiodic components. Nat.
Neurosci. 23 (12), 1655U288.
Doppelmayr, M., Klimesch, W., Pachinger, T., Ripper, B., 1998. Individual differences in
brain dynamics: important implications for the calculation of event-related band
power. Biol. Cybern. 79 (1), 4957.
Dotson, V.M., Duarte, A., 2020. The importance of diversity in cognitive neuroscience.
Ann. N. Y. Acad. Sci. 1464 (1), 181191.
Elad, D., Cetin-Karayumak, S., Zhang, F., Cho, K.I.K., Lyall, A.E., Seitz-Holland, J., et al.,
2021. Improving the predictive potential of diffusion MRI in schizophrenia using
normative models-Towards subject-level classication. Hum. Brain Mapp. 42 (14),
46584670.
Faul, F., Erdfelder, E., Lang, A.G., G*Power, Buchner A., 2007. 3: a exible statistical
power analysis program for the social, behavioral, and biomedical sciences. Behav.
Res. Methods 39 (2), 175191.
Finnigan, S., Robertson, I.H., 2011. Resting EEG theta power correlates with cognitive
performance in healthy older adults. Psychophysiology. 48 (8), 10831087.
Fisher, R.A., 1992. Statistical Methods for Research Workers. In: Kotz, S, Johnson, NL
(Eds.), Breakthroughs in Statistics: Methodology and Distribution. Springer New
York, New York, NY, pp. 6670.
Fittipaldi, S., Legaz, A., Maito, M., Hernandez, H., Altschuler, F., Canziani, V., et al.,
2023. Heterogeneous factors inuence social cognition across diverse settings in
brain health and age-related diseases. Res. Sq.
Foderaro, G., Isella, V., Mazzone, A., Biglia, E., Di Gangi, M., Pasotti, F., et al., 2022.
Brand new norms for a good old test: northern Italy normative study of MiniMental
State Examination. Neurol. Sci. 43 (5), 30533063.
Folstein, M.F., Folstein, S.E., McHugh, P.R., 1975. "Mini-mental state". A practical
method for grading the cognitive state of patients for the clinician. J. Psychiatr. Res.
12 (3), 189198.
Franzmeier, N., Duering, M., Weiner, M., Dichgans, M., Ewers, M., 2017b. Alzheimers
Disease Neuroimaging I. Left frontal cortex connectivity underlies cognitive reserve
in prodromal Alzheimer disease. Neurology. 88 (11), 10541061.
Franzmeier, N., Gottler, J., Grimmer, T., Drzezga, A., Araque-Caballero, M.A., Simon-
Vermot, L., et al., 2017a. Resting-state connectivity of the left frontal cortex to the
default mode and dorsal attention network supports reserve in mild cognitive
impairment. Front. Aging Neurosci. 9, 264.
Gaal, Z.A., Boha, R., Stam, C.J., Molnar, M., 2010. Age-dependent features of EEG-
reactivityspectral, complexity, and network characteristics. Neurosci. Lett. 479 (1),
7984.
Garcia, M.A., Saenz, J., Downer, B., Wong, R., 2018. The role of education in the
association between race/ethnicity/nativity, cognitive impairment, and dementia
among older adults in the United States. Demogr. Res. 38, 155168.
H. Hernandez et al.
NeuroImage 295 (2024) 120636
18
Garo-Pascual, M., Gaser, C., Zhang, L., Tohka, J., Medina, M., Strange, B.A., 2023. Brain
structure and phenotypic prole of superagers compared with age-matched older
adults: a longitudinal analysis from the Vallecas Project. Lancet Healthy. Longev. 4
(8), e374ee85.
Gaubert, S., Raimondo, F., Houot, M., Corsi, M.C., Naccache, L., Diego, Sitt J, et al., 2019.
EEG evidence of compensatory mechanisms in preclinical Alzheimers disease. Brain
142 (7), 20962112.
Gordon, B.A., Rykhlevskaia, E.I., Brumback, C.R., Lee, Y., Elavsky, S., Konopack, J.F.,
et al., 2008. Neuroanatomical correlates of aging, cardiopulmonary tness level, and
education. Psychophysiology. 45 (5), 825838.
G´
orecka, J., Walerjan, P., 2011. Artifacts Extraction from EEG Data Using the Infomax
Approach. Biocybern. Biomed. Eng. 31, 5974.
Grech, R., Cassar, T., Muscat, J., Camilleri, K.P., Fabri, S.G., Zervakis, M., et al., 2008.
Review on solving the inverse problem in EEG source analysis. J. Neuroeng. Rehabil.
5, 25.
Greene, A.S., Gao, S., Noble, S., Scheinost, D., Constable, R.T., 2020. How Tasks change
whole-brain functional organization to reveal brain-phenotype relationships. Cell
Rep. 32 (8), 108066.
Greene, A.S., Shen, X., Noble, S., Horien, C., Hahn, C.A., Arora, J., et al., 2022. Brain-
phenotype models fail for individuals who defy sample stereotypes. Nature 609
(7925), 109118.
Habes, M., Pomponio, R., Shou, H., Doshi, J., Mamourian, E., Erus, G., et al., 2021. The
Brain Chart of Aging: machine-learning analytics reveals links between brain aging,
white matter disease, amyloid burden, and cognition in the iSTAGING consortium of
10,216 harmonized MR scans. Alzheimers. Dement. 17 (1), 89102.
Haegens, S., Cousijn, H., Wallis, G., Harrison, P.J., Nobre, A.C, 2014. Inter- and intra-
individual variability in alpha peak frequency. Neuroimage 92 (100), 4655.
Heger, I.S., Deckers, K., Schram, M.T., Stehouwer, C.D.A., Dagnelie, P.C., van der
Kallen, C.J.H., et al., 2021. Associations of the Lifestyle for Brain Health Index With
Structural Brain Changes and Cognition Results From the Maastricht Study.
Neurology. 97 (13), E1300E1E12.
Hemmati, S., Ahmadlou, M., Gharib, M., Vameghi, R., Sajedi, F., 2013. Down syndromes
brain dynamics: analysis of fractality in resting state. Cogn. Neurodyn. 7 (4),
333340.
Herzog, R., Rosas, F.E., Whelan, R., Fittipaldi, S., Santamaria-Garcia, H., Cruzat, J., et al.,
2022. Genuine high-order interactions in brain networks and neurodegeneration.
Neurobiol. Dis. 175, 105918.
Hill, A.T., Bailey, N.W., Zomorrodi, R., Hadas, I., Kirkovski, M., Das, S., et al., 2023. EEG
microstates in early-to-middle childhood show associations with age, biological sex,
and alpha power. Hum. Brain Mapp. 44 (18), 64846498.
Hill, A.T., Clark, G.M., Bigelow, F.J., Lum, J.A.G., Enticott, P.G., 2022. Periodic and
aperiodic neural activity displays age-dependent changes across early-to-middle
childhood. Dev. Cogn. Neurosci. 54, 101076.
Hinault, T., Baillet, S., Courtney, S.M., 2023. Age-related changes of deep-brain
neurophysiological activity. Cereb. Cortex. 33 (7), 39603968.
Hindriks, R., van Putten, M.J., 2013. Thalamo-cortical mechanisms underlying changes
in amplitude and frequency of human alpha oscillations. Neuroimage 70, 150163.
Holmes, A.J., Patrick, L.M., 2018. The myth of optimality in clinical neuroscience.
Trends. Cogn. Sci. 22 (3), 241257.
Hu, S., Lai, Y., Valdes-Sosa, P.A., Bringas-Vega, M.L., Yao, D., 2018. How do reference
montage and electrodes setup affect the measured scalp EEG potentials? J. Neural
Eng. 15 (2), 026013.
Hunt, B.A.E., Wong, S.M., Vandewouw, M.M., Brookes, M.J., Dunkley, B.T., Taylor, M.J.,
2019. Spatial and spectral trajectories in typical neurodevelopment from childhood
to middle age. Netw. Neurosci. 3 (2), 497520.
Ibanez, A., Legaz, A., Ruiz-Adame, M., 2023. Addressing the gaps between
socioeconomic disparities and biological models of dementia. Brain 146, 35613564.
Iinuma, Y., Nobukawa, S., Mizukami, K., Kawaguchi, M., Higashima, M., Tanaka, Y.,
et al., 2022. Enhanced temporal complexity of EEG signals in older individuals with
high cognitive functions. Front. Neurosci. 16, 878495.
Imperatori, L.S., Betta, M., Cecchetti, L., Canales-Johnson, A., Ricciardi, E., Siclari, F.,
et al., 2019. EEG functional connectivity metrics wPLI and wSMI account for distinct
types of brain functional interactions. Sci. Rep. 9 (1), 8894.
Ince, R.A., Giordano, B.L., Kayser, C., Rousselet, G.A., Gross, J., Schyns, P.G., 2017.
A statistical framework for neuroimaging data analysis based on mutual information
estimated via a gaussian copula. Hum. Brain Mapp. 38 (3), 15411573.
Ishii, R., Canuet, L., Aoki, Y., Hata, M., Iwase, M., Ikeda, S., et al., 2017. Healthy and
pathological brain aging: from the perspective of oscillations, functional
connectivity, and signal complexity. Neuropsychobiology. 75 (4), 151161.
Javaid, H., Kumarnsit, E., Chatpun, S., 2022. Age-related alterations in EEG network
connectivity in healthy aging. Brain Sci. 12 (2).
Jawinski, P., Markett, S., Drewelies, J., Duzel, S., Demuth, I., Steinhagen-Thiessen, E.,
et al., 2022. Linking Brain Age Gap to Mental and Physical Health in the Berlin Aging
Study II. Front. Aging Neurosci. 14, 791222.
Jeong, H.T., Youn, Y.C., Sung, H.H., Kim, S.Y., 2021. Power spectral changes of
quantitative eeg in the subjective cognitive decline: comparison of community
normal control groups. Neuropsychiatr. Dis. Treat. 17, 27832790.
Jokinen, H., Melkas, S., Madureira, S., Verdelho, A., Ferro, J.M., Fazekas, F., et al., 2016.
Cognitive reserve moderates long-term cognitive and functional outcome in cerebral
small vessel disease. J. Neurol. Neurosurg. Psychiatry 87 (12), 12961302.
Jovicich, J., Barkhof, F., Babiloni, C., Herholz, K., Mulert, C., 2019. Harmonization of
Neuroimaging Biomarkers For Neurodegenerative diseases: A survey in the Imaging
Community of Perceived Barriers and Suggested Actions, 11. Alzheimers &
dementia, (Amsterdam, Netherlands), pp. 6973.
Kalauzi, A., Boji´
c, T., Raki´
c, L., 2009. Extracting complexity waveforms from one-
dimensional signals. Nonlinear. Biomed. Phys. 3 (1), 8.
Kida, T., Tanaka, E., Kakigi, R., 2016. Multi-Dimensional Dynamics of Human
Electromagnetic Brain Activity. Front. Hum. Neurosci. 9.
Klimesch, W., 1999b. EEG alpha and theta oscillations reect cognitive and memory
performance: a review and analysis. Brain Res. Brain Res. Rev. 29 (23), 169195.
Klimesch, W., 1999a. EEG alpha and theta oscillations reect cognitive and memory
performance: a review and analysis. Brain Res. Rev. 29 (2), 169195.
Kluger, D.S., Forster, C., Abbasi, O., Chalas, N., Villringer, A., Gross, J., 2023.
Modulatory dynamics of periodic and aperiodic activity in respiration-brain
coupling. Nat. Commun. 14 (1), 4699.
Kochhann, R., Varela, J.S., Lisboa, C.S.M., Chaves, M.L.F., 2010. The Mini Mental State
Examination: review of cutoff points adjusted for schooling in a large Southern
Brazilian sample. Dement. Neuropsychol. 4 (1), 3541.
Koenig, T., Prichep, L., Lehmann, D., Sosa, P.V., Braeker, E., Kleinlogel, H., et al., 2002.
Millisecond by millisecond, year by year: normative EEG microstates and
developmental stages. Neuroimage 16 (1), 4148.
Kothe, C.A., Makeig, S., 2013. BCILAB: a platform for brain-computer interface
development. J. Neural Eng. 10 (5), 056014.
Krause, D., Folkerts, M., Karch, S., Keeser, D., Chrobok, A.I., Zaudig, M., et al., 2015.
Prediction of Treatment outcome in patients with obsessive-compulsive disorder
with low-resolution brain electromagnetic tomography: a prospective EEG study.
Front. Psychol. 6, 1993.
Lau, Z.J., Pham, T., Chen, S.H.A., Makowski, D., 2022. Brain entropy, fractal dimensions
and predictability: a review of complexity measures for EEG in healthy and
neuropsychiatric populations. Eur. J. Neurosci. 56 (7), 50475069.
Lee, H.Y., Jung, K.I., Yoo, W.K., Ohn, S.H., 2019. Global Synchronization Index as an
Indicator for Tracking Cognitive Function Changes in a Traumatic Brain Injury
Patient: a Case Report. Ann. Rehabil. Med. 43 (1), 106110.
Lejko, N., Larabi, D.I., Herrmann, C.S., Aleman, A., Curcic-Blake, B., 2020. Alpha power
and functional connectivity in cognitive decline: a systematic review and meta-
analysis. J. Alzheimers. Dis. 78 (3), 10471088.
Li, X., Ouyang, G., Richards, D.A., 2007. Predictability analysis of absence seizures with
permutation entropy. Epilepsy Res. 77 (1), 7074.
Liao, X., Vasilakos, A.V., He, Y., 2017. Small-world human brain networks: perspectives
and challenges. Neurosci. Biobehav. Rev. 77, 286300.
Livingston, G., Huntley, J., Sommerlad, A., Ames, D., Ballard, C., Banerjee, S., et al.,
2020. Dementia prevention, intervention, and care: 2020 report of the Lancet
Commission. Lancet. 396 (10248), 413446.
Manly, B.F.J, 1997. Randomization, Bootstrap and Monte Carlo Methods in Biology,
second edition. Taylor & Francis.
Marek, S., Tervo-Clemmens, B., Calabro, F.J., Montez, D.F., Kay, B.P., Hatoum, A.S.,
et al., 2022. Reproducible brain-wide association studies require thousands of
individuals. Nature 603 (7902), 654660.
Marquand, A.F., Rezek, I., Buitelaar, J., Beckmann, C.F., 2016. Understanding
heterogeneity in clinical cohorts using normative models: beyond case-control
studies. Biol. Psychiatry 80 (7), 552561.
Martinez-Canada, P., Perez-Valero, E., Minguillon, J., Pelayo, F., Lopez-Gordo, M.A.,
Morillas, C., 2023. Combining aperiodic 1/f slopes and brain simulation: an EEG/
MEG proxy marker of excitation/inhibition imbalance in Alzheimers disease.
Alzheimers. Dement. (Amst) 15 (3), e12477.
Matshabane, O.P., 2021. Promoting diversity and inclusion in neuroscience and
neuroethics. EBioMedicine 67, 103359.
Maturana-Candelas, A., Gomez, C., Poza, J., Pinto, N., Hornero, R., 2019. EEG
Characterization of the Alzheimers Disease Continuum by Means of Multiscale
Entropies. Entropy. (Basel) 21 (6).
Mazziotta, J., Toga, A., Evans, A., Fox, P., Lancaster, J., Zilles, K., et al., 2001.
A probabilistic atlas and reference system for the human brain: international
Consortium for Brain Mapping (ICBM). Philos. Trans. R. Soc. Lond. B Biol. Sci. 356
(1412), 12931322.
McBride, J.C., Zhao, X., Munro, N.B., Smith, C.D., Jicha, G.A., Hively, L., et al., 2014.
Spectral and complexity analysis of scalp EEG characteristics for mild cognitive
impairment and early Alzheimers disease. Comput. Methods Programs Biomed. 114
(2), 153163.
Medel, V., Irani, M., Crossley, N., Ossandon, T., Boncompte, G., 2023b. Complexity and
1/f slope jointly reect brain states. Sci. Rep. 13 (1), 21700.
Medel, V., Irani, M., Crossley, N., Ossand´
on, T., Boncompte, G., 2023a. Complexity and
1/f slope jointly reect brain states. Sci. Rep. 13 (1), 21700.
Melnik, A., Legkov, P., Izdebski, K., K¨
archer, S.M., Hairston, W.D., Ferris, D.P., et al.,
2017. Systems, Subjects, Sessions: to What Extent Do These Factors Inuence EEG
Data? Front. Hum. Neurosci. 11, 150.
Members, E.C.C., Brayne, C., Ince, P.G., Keage, H.A., McKeith, I.G., Matthews, F.E., et al.,
2010. Education, the brain and dementia: neuroprotection or compensation? Brain
133 (Pt 8), 22102216.
Merkin, A., Sghirripa, S., Graetz, L., Smith, A.E., Hordacre, B., Harris, R., et al., 2023. Do
age-related differences in aperiodic neural activity explain differences in resting EEG
alpha? Neurobiol. Aging 121, 7887.
Meunier, D., Achard, S., Morcom, A., Bullmore, E., 2009. Age-related changes in modular
organization of human brain functional networks. Neuroimage 44 (3), 715723.
Miraglia, F., Vecchio, F., Bramanti, P., Rossini, P.M., 2015. Small-worldness
characteristics and its gender relation in specic hemispheric networks.
Neuroscience 310, 111.
Moretti, D.V., Babiloni, C., Binetti, G., Cassetta, E., Dal Forno, G., Ferreric, F., et al.,
2004b. Individual analysis of EEG frequency and band power in mild Alzheimers
disease. Clin. Neurophysiol. 115 (2), 299308.
Moretti, D.V., Babiloni, C., Binetti, G., Cassetta, E., Dal Forno, G., Ferreric, F., et al.,
2004a. Individual analysis of EEG frequency and band power in mild Alzheimers
disease. Clin. Neurophysiol. 115 (2), 299308.
H. Hernandez et al.
NeuroImage 295 (2024) 120636
19
Mukaetova-Ladinska, E.B., De Lillo, C., Arshad, Q., Subramaniam, H.E., Maltby, J., 2022.
Cognitive Assessment of Dementia: the Need for an Inclusive Design Tool. Curr.
Alzheimer. Res. 19 (4), 265273.
Muller, A.C., Guido, S., 2018. Introduction to Machine Learning with Python: a Guide for
Data Scientists: oReilly Media. Incorporated.
Murty, D., Manikandan, K., Kumar, W.S., Ramesh, R.G., Purokayastha, S., Javali, M.,
et al., 2020. Gamma oscillations weaken with age in healthy elderly in human EEG.
Neuroimage 215, 116826.
Nguyen-Danse, D.A., Singaravelu, S., Chauvign´
e, L.A.S., Mottaz, A., Allaman, L.,
Guggisberg, A.G., 2021. Feasibility of reconstructing source functional connectivity
with low-density EEG. Brain Topogr. 34 (6), 709719.
Onoda, K., Yamaguchi, S., 2013. Small-worldness and modularity of the resting-state
functional brain network decrease with aging. Neurosci. Lett. 556, 104108.
Ouyang, C.S., Chiang, C.T., Yang, R.C., Wu, R.C., Lin, L.C., 2020b. Quantitative
electroencephalogram analysis of frontal cortex functional changes in patients with
migraine. Kaohsiung. J. Med. Sci. 36 (7), 543551.
Ouyang, G., Hildebrandt, A., Schmitz, F., Herrmann, C.S., 2020a. Decomposing alpha
and 1/f brain activities reveals their differential associations with cognitive
processing speed. Neuroimage 205, 116304.
Ouyang, G., Li, J., Liu, X., Li, X., 2013. Dynamic characteristics of absence EEG
recordings with multiscale permutation entropy analysis. Epilepsy Res. 104 (3),
246252.
Parbat, D., Chakraborty, M., 2021. A novel methodology to study the cognitive load
induced eeg complexity changes: chaos, fractal and entropy based approach. Biomed
Signal Proces 64.
Parra, M.A., 2022. Barriers to effective memory assessments for Alzheimers Disease.
J. Alzheimers. Dis. 90 (3), 981988.
Pascual-Marqui, R.D., 2002. Standardized low-resolution brain electromagnetic
tomography (sLORETA): technical details. Methods Find. Exp. Clin. Pharmacol. 24
(Suppl D), 512.
Pathania, A., Schreiber, M., Miller, M.W., Euler, M.J., Lohse, K.R., 2021. Exploring the
reliability and sensitivity of the EEG power spectrum as a biomarker. Int. J.
Psychophysiol. 160, 1827.
Pei, L., Zhou, X., Leung, F.K.S., Ouyang, G., 2023. Differential associations between
scale-free neural dynamics and different levels of cognitive ability.
Psychophysiology. 60 (6), e14259.
Pfurtscheller, G., Lopes da Silva, F.H., 1999. Event-related EEG/MEG synchronization
and desynchronization: basic principles. Clin. Neurophysiol. 110 (11), 18421857.
Pinaya, W.H.L., Scarpazza, C., Garcia-Dias, R., Vieira, S., Baecker, L., FdC, P., et al., 2021.
Using normative modelling to detect disease progression in mild cognitive
impairment and Alzheimers disease in a cross-sectional multi-cohort study. Sci. Rep.
11 (1), 15746.
Pion-Tonachini, L., Kreutz-Delgado, K., Makeig, S., 2019. ICLabel: an automated
electroencephalographic independent component classier, dataset, and website.
Neuroimage 198, 181197.
Prado, P., Birba, A., Cruzat, J., Santamaria-Garcia, H., Parra, M., Moguilner, S., et al.,
2022. Dementia ConnEEGtome: towards multicentric harmonization of EEG
connectivity in neurodegeneration. Int. J. Psychophysiol. 172, 2438.
Prado, P., Mejia, J.A., Sainz-Ballesteros, A., Birba, A., Moguilner, S., Herzog, R., et al.,
2023a. Harmonized multi-metric and multi-centric assessment of EEG source space
connectivity for dementia characterization. Alzheimers. Dement. (Amst) 15 (3),
e12455.
Prado, P., Mejía, J.A., Sainz-Ballesteros, A., Birba, A., Moguilner, S., Herzog, R., et al.,
2023c. Harmonized multi-metric and multi-centric assessment of EEG source space
connectivity for dementia characterization. Alzheimers. Dement. (Amst) 15 (3),
e12455.
Prado, P., Moguilner, S., Mejia, J.A., Sainz-Ballesteros, A., Otero, M., Birba, A., et al.,
2023b. Source space connectomics of neurodegeneration: one-metric approach does
not t all. Neurobiol. Dis. 179, 106047.
Pravitha, R., Sreenivasan, R., Nampoori, V.P., 2005. Complexity analysis of dense array
EEG signal reveals sex difference. Int. J. Neurosci. 115 (4), 445460.
Raz, N., Lindenberger, U., Rodrigue, K.M., Kennedy, K.M., Head, D., Williamson, A.,
et al., 2005. Regional brain changes in aging healthy adults: general trends,
individual differences and modiers. Cereb. Cortex. 15 (11), 16761689.
Rempe, M.P., Ott, L.R., Picci, G., Penhale, S.H., Christopher-Hayes, N.J., Lew, B.J., et al.,
2023. Spontaneous cortical dynamics from the rst years to the golden years. Proc
Natl Acad Sci U S A. 120 (4), e2212776120.
Resende, E.P.F., Llibre, Guerra JJ, Miller, B.L, 2019. Health and Socioeconomic
Inequities as Contributors to Brain Health. JAMa Neurol. 76 (6), 633634.
Rolls, E.T., Joliot, M., Tzourio-Mazoyer, N., 2015. Implementation of a new parcellation
of the orbitofrontal cortex in the automated anatomical labeling atlas. Neuroimage
122, 15.
Rosenberg, M.D., Scheinost, D., Greene, A.S., Avery, E.W., Kwon, Y.H., Finn, E.S., et al.,
2020. Functional connectivity predicts changes in attention observed across minutes,
days, and months. Proc. Natl. Acad. Sci. U. S. A 117 (7), 37973807.
Rossini, P.M., Di Iorio, R., Vecchio, F., Anfossi, M., Babiloni, C., Bozzali, M., et al., 2020.
Early diagnosis of Alzheimers disease: the role of biomarkers including advanced
EEG signal analysis. Report from the IFCN-sponsored panel of experts. Clin.
Neurophysiol. 131 (6), 12871310.
Rossini, P.M., Rossi, S., Babiloni, C., Polich, J., 2007. Clinical neurophysiology of aging
brain: from normal aging to neurodegeneration. Prog. Neurobiol. 83 (6), 375400.
Rubinov, M., Sporns, O., 2010. Complex network measures of brain connectivity: uses
and interpretations. Neuroimage 52 (3), 10591069.
Rutherford, S., Fraza, C., Dinga, R., Kia, S.M., Wolfers, T., Zabihi, M., et al., 2022.
Charting brain growth and aging at high spatial precision. Elife 11.
Santamaria-Garcia, H., Sainz-Ballesteros, A., Hernandez, H., Moguilner, S., Maito, M.,
Ochoa-Rosales, C., et al., 2023. Factors associated with healthy aging in Latin
American populations. Nat. Med. 29 (9), 22482258.
Sarasso, S., Casali, A.G., Casarotto, S., Rosanova, M., Sinigaglia, C., Massimini, M., 2021.
Consciousness and complexity: a consilience of evidence. Neuroscience of
Consciousness niab023.
Schliebs, R., Arendt, T., 2011. The cholinergic system in aging and neuronal
degeneration. Behav. Brain Res. 221 (2), 555563.
Schreckenberger, M., Lange-Asschenfeldt, C., Lochmann, M., Mann, K., Siessmeier, T.,
Buchholz, H.G., et al., 2004. The thalamus as the generator and modulator of EEG
alpha rhythm: a combined PET/EEG study with lorazepam challenge in humans.
Neuroimage 22 (2), 637644.
Seker, M., Ozbek, Y., Yener, G., Ozerdem, M.S., 2021. Complexity of EEG Dynamics for
Early Diagnosis of Alzheimers Disease Using Permutation Entropy Neuromarker.
Comput. Methods Programs Biomed. 206, 106116.
Selya, A.S., Rose, J.S., Dierker, L.C., Hedeker, D., Mermelstein, R.J., 2012. A practical
guide to calculating Cohens f(2), a Measure of Local Effect Size, from PROC MIXED.
Front. Psychol. 3, 111.
Senner, N.R., Conklin, J.R., Piersma, T., 1814. An ontogenetic perspective on individual
differences. Proc. Biol. Sci. 282, 2015.
Shumbayawonda, E., Tosun, P.D., Fernandez, A., Hughes, M.P., Abasolo, D., 2018.
Complexity changes in brain activity in healthy ageing: a permutation lempel-ziv
complexity study of magnetoencephalograms. Entropy. (Basel) 20 (7).
Singh, N.C., 2011. Measuring the ‘complexityof sound. Pramana 77 (5), 811816.
Smith, A.E., Chau, A., Greaves, D., Keage, H.A.D., Feuerriegel, D., 2023. Resting EEG
power spectra across middle to late life: associations with age, cognition, APOE-
varepsilon4 carriage, and cardiometabolic burden. Neurobiol. Aging 130, 93102.
Smits, F.M., Porcaro, C., Cottone, C., Cancelli, A., Rossini, P.M., Tecchio, F., 2016.
Electroencephalographic Fractal Dimension in Healthy Ageing and Alzheimers
Disease. PLoS. One 11 (2), e0149587.
Snee, R., 1983. Regression diagnostics: identifying inuential data and sources of
collinearity. J. Qual. Technol. 15, 149153.
Soler, A., Mu˜
noz-Guti´
errez, P.A., Bueno-L´
opez, M., Giraldo, E., Molinas, M, 2020. Low-
Density EEG for neural activity reconstruction using multivariate empirical mode
decomposition. Front. Neurosci. 14, 175.
Song, J., Birn, R.M., Boly, M., Meier, T.B., Nair, V.A., Meyerand, M.E., et al., 2014. Age-
related reorganizational changes in modularity and functional connectivity of
human brain networks. Brain Connect. 4 (9), 662676.
Stacey, J.E., Crook-Rumsey, M., Sumich, A., Howard, C.J., Crawford, T., Livne, K., et al.,
2021. Age differences in resting state EEG and their relation to eye movements and
cognitive performance. Neuropsychologia 157, 107887.
Stanley, M.L., Simpson, S.L., Dagenbach, D., Lyday, R.G., Burdette, J.H., Laurienti, P.J.,
2015. Changes in brain network efciency and working memory performance in
aging. PLoS. One 10 (4), e0123950.
Stier, C., Braun, C., Focke, N.K., 2023. Adult lifespan trajectories of neuromagnetic
signals and interrelations with cortical thickness. Neuroimage 278, 120275.
Sun, J., Wang, B., Niu, Y., Tan, Y., Fan, C., Zhang, N., et al., 2020. Complexity Analysis of
EEG, MEG, and fMRI in Mild Cognitive Impairment and Alzheimers Disease: a
Review. Entropy. (Basel) 22 (2).
Tan, Y., Chen, J., Liao, W., Qian, Z., 2019. Brain function network and young adult
smokers: a graph theory analysis study. Front. Psychiatry 10, 590.
Tang, A.S., Oskotsky, T., Havaldar, S., Mantyh, W.G., Bicak, M., Solsberg, C.W., et al.,
2022. Deep phenotyping of Alzheimers disease leveraging electronic medical
records identies sex-specic clinical associations. Nat. Commun. 13 (1), 675.
Tomescu, M.I., Rihs, T.A., Rochas, V., Hardmeier, M., Britz, J., Allali, G., et al., 2018.
From swing to cane: sex differences of EEG resting-state temporal patterns during
maturation and aging. Dev. Cogn. Neurosci. 31, 5866.
Tononi, G., Sporns, O., Edelman, G.M., 1994. A measure for brain complexity: relating
functional segregation and integration in the nervous system. Proc. Natl. Acad. Sci.
91 (11), 50335037.
Trammell, J.P., MacRae, P.G., Davis, G., Bergstedt, D., Anderson, A.E., 2017. The
relationship of cognitive performance and the theta-alpha power ratio is age-
dependent: an eeg study of short term memory and reasoning during task and
resting-state in healthy young and old adults. Front. Aging Neurosci. 9, 364.
Trondle, M., Popov, T., Pedroni, A., Pfeiffer, C., Baranczuk-Turska, Z., Langer, N., 2023.
Decomposing age effects in EEG alpha power. Cortex 161, 116144.
Tsigler, A., Bartlett, P.L., 2024. Benign overtting in ridge regression. J. Mach. Learn.
Res. 24 (1), 123.
Turner, C., Baylan, S., Bracco, M., Cruz, G., Hanzal, S., Keime, M., et al., 2023.
Developmental changes in individual alpha frequency: recording EEG data during
public engagement events. Imaging Neurosci (Camb) 1, 114.
Valdes-Sosa, P.A., Galan-Garcia, L., Bosch-Bayard, J., Bringas-Vega, M.L., Aubert-
Vazquez, E., Rodriguez-Gil, I., et al., 2021. The Cuban Human Brain Mapping
Project, a young and middle age population-based EEG, MRI, and cognition dataset.
Sci. Data 8 (1), 45.
Valsd´
ottir, V., Magnúsd´
ottir, B.B., Chang, M.L., Sigurdsson, S., Gudnason, V., Launer, L.
J., et al., 2022. Cognition and brain health among older adults in Iceland: the AGES-
Reykjavik study. Geroscience 44 (6), 27852800.
van Nifterick, A.M., Mulder, D., Duineveld, D.J., Diachenko, M., Scheltens, P., Stam, C.J.,
et al., 2023. Resting-state oscillations reveal disturbed excitation-inhibition ratio in
Alzheimers disease patients. Sci. Rep. 13 (1), 7419.
Varela, F., Lachaux, J.P., Rodriguez, E., Martinerie, J., 2001. The brainweb: phase
synchronization and large-scale integration. Nat. Rev. Neurosci. 2 (4), 229239.
Vecchio, F., Miraglia, F., Alu, F., Judica, E., Cotelli, M., Pellicciari, M.C., et al., 2022.
Human brain networks in physiological and pathological aging: reproducibility of
H. Hernandez et al.
NeuroImage 295 (2024) 120636
20
electroencephalogram graph theoretical analysis in cortical connectivity. Brain
Connect. 12 (1), 4151.
Verdi, S., Kia, S.M., Schott, J.M., Marquand, A., Cole, J.H., 2022. Mapping individualised
patterns of atrophy in Alzheimers disease using neuroanatomical normative models.
Alzheimers. Dement. 18 (56).
Voytek, B., Kramer, M.A., Case, J., Lepage, K.Q., Tempesta, Z.R., Knight, R.T., et al.,
2015. Age-Related Changes in 1/f Neural Electrophysiological Noise. J. Neurosci. 35
(38), 1325713265.
Walsh, S., Merrick, R., Brayne, C., 2022. The relevance of social and commercial
determinants for neurological health. Lancet Neurol. 21 (12), 11511160.
Wang, H., McIntosh, A.R., Kovacevic, N., Karachalios, M., Protzner, A.B., 2016. Age-
related Multiscale Changes in Brain Signal Variability in Pre-task versus Post-task
Resting-state EEG. J. Cogn. Neurosci. 28 (7), 971984.
Wang, R., Wang, J., Yu, H., Wei, X., Yang, C., Deng, B, 2015. Power spectral density and
coherence analysis of Alzheimers EEG. Cogn. Neurodyn. 9 (3), 291304.
Waschke, L., Donoghue, T., Fiedler, L., Smith, S., Garrett, D.D., Voytek, B., et al., 2021.
Modality-specic tracking of attention and sensory statistics in the human
electrophysiological spectral exponent. Elife 10.
Wilkinson, C.L., Pierce, L.J., Sideridis, G., Wade, M., Nelson, C.A., 2023. Associations
between EEG trajectories, family income, and cognitive abilities over the rst two
years of life. Dev. Cogn. Neurosci. 61, 101260.
Wolfers, T., Beckmann, C.F., Hoogman, M., Buitelaar, J.K., Franke, B., Marquand, A.F.,
2020. Individual differences v. the average patient: mapping the heterogeneity in
ADHD using normative models. Psychol. Med. 50 (2), 314323.
Yan, J., Li, T., Wang, H., Huang, H., Wan, J., Nho, K., et al., 2015. Cortical surface
biomarkers for predicting cognitive outcomes using group l2,1 norm. Neurobiol.
Aging 36 (Suppl 1), S185S193.
Yu, M., Sporns, O., Saykin, A.J., 2021. Author Correction: the human connectome in
Alzheimer disease - relationship to biomarkers and genetics. Nat. Rev. Neurol. 17
(9), 592.
Zappasodi, F., Marzetti, L., Olejarczyk, E., Tecchio, F., Pizzella, V., 2015. Age-related
changes in electroencephalographic signal complexity. PLoS One. 10 (11),
e0141995.
Zhang, C., Stock, A.K., Muckschel, M., Hommel, B., Beste, C., 2023. Aperiodic neural
activity reects metacontrol. Cereb. Cortex. 33 (12), 79417951.
Zhao, L., Zhang, Y., Yu, X., Wu, H., Wang, L., Li, F., et al., 2023. Quantitative signal
quality assessment for large-scale continuous scalp electroencephalography from a
big data perspective. Physiol. Meas. 44 (3).
H. Hernandez et al.
... Brain age indices has been restricted by the predominant use of MRI or PET, which are less accessible and affordable in LAC, leading to selection biases 34 . EEG offers a solution due to its cost-effectiveness, portability, and ease of implementation in aging and dementia 35,36 . However, few studies have combined accessible techniques with deep learning to develop scalable brain age markers. ...
... Our framework combines multiple dimensions of diversity in brain health, the Alzheimer's disease continuum and related disorders within a single measure of brain clocks, which is relevant for global health policies, generalizable computational models, and public health strategies. Incorporating EEG offers affordable and scalable solutions for disadvantaged settings, such as those in LAC, compared to traditional neuroimaging techniques 1,35 . Accessible metrics of accelerated aging can offer personalized assessments of diversity, aging, and neurocognitive disorders. ...
Preprint
Full-text available
Brain clocks, which quantify discrepancies between brain age and chronological age, hold promise for understanding brain health and disease. However, the impact of multimodal diversity (geographical, socioeconomic, sociodemographic, sex, neurodegeneration) on the brain age gap (BAG) is unknown. Here, we analyzed datasets from 5,306 participants across 15 countries (7 Latin American countries -LAC, 8 non-LAC). Based on higher-order interactions in brain signals, we developed a BAG deep learning architecture for functional magnetic resonance imaging (fMRI=2,953) and electroencephalography (EEG=2,353). The datasets comprised healthy controls, and individuals with mild cognitive impairment, Alzheimer’s disease, and behavioral variant frontotemporal dementia. LAC models evidenced older brain ages (fMRI: MDE=5.60, RMSE=11.91; EEG: MDE=5.34, RMSE=9.82) compared to non-LAC, associated with frontoposterior networks. Structural socioeconomic inequality and other disparity-related factors (pollution, health disparities) were influential predictors of increased brain age gaps, especially in LAC (R²=0.37, F²=0.59, RMSE=6.9). A gradient of increasing BAG from controls to mild cognitive impairment to Alzheimer’s disease was found. In LAC, we observed larger BAGs in females in control and Alzheimer’s disease groups compared to respective males. Results were not explained by variations in signal quality, demographics, or acquisition methods. Findings provide a quantitative framework capturing the multimodal diversity of accelerated brain aging.
... Additionally, the cognitive traits were reported to be different between genders; for example, navigation by mental maps, mental rotation, visuo-spatial skills, and working memory were stronger in men, while navigation by landmarks, verbal ability, reading and writing abilities, fine motor coordination, and perceptual skills higher in women [1]. Brain health in diverse settings, including geographical differences, age, cognition, and demographics, can affect EEG-related changes [64][65][66][67]. This study provides insights on the neuropsychological test profile between genders in a sample of the Turkish population in the West Asia region. ...
Article
Full-text available
Earlier research has suggested gender differences in event-related potentials/oscillations (ERPs/EROs). Yet, the alteration in event-related oscillations (EROs) in the delta and theta frequency bands have not been explored between genders across the three age groups of adulthood, i.e., 18–50, 51–65, and >65 years. Data from 155 healthy elderly participants who underwent a neurological examination, comprehensive neuropsychological assessment (including attention, memory, executive function, language, and visuospatial skills), and magnetic resonance imaging (MRI) from past studies were used. The delta and theta ERO powers across the age groups and between genders were compared and correlational analyses among the ERO power, age, and neuropsychological tests were performed. The results indicated that females displayed higher theta ERO responses than males in the frontal, central, and parietal regions but not in the occipital location between 18 and 50 years of adulthood. The declining theta power of EROs in women reached that of men after the age of 50 while the theta ERO power was more stable across the age groups in men. Our results imply that the cohorts must be recruited at specified age ranges across genders, and clinical trials using neurophysiological biomarkers as an intervention endpoint should take gender into account in the future.
Preprint
Full-text available
Aging affects brain structure and function alongside metabolic and vascular processes leading to energetic impairments. While local neurometabolic dysfunction in aging is well-documented, the influence of systemic cardiometabolic and vascular markers on brain structure and function remains less understood. We examine the link between cardiometabolic dysfunction (measured by an allostatic load index) and neurovascular burden (measured by white matter hyperintensities) with brain changes, including ventricular and hippocampal volume, as well as EEG activity, across age. Analyzing data from 196 healthy individuals across age (20-75 years), we found a significant positive correlation between allostatic load index and white-matter hyperintensities, irrespective of age. White matter hyperintensities are also positively linked with ventricular enlargement, but not hippocampal atrophy. The allostatic load index mediated the relationship between white-matter hyperintensities and ventricular volume. Regarding brain function, changes in the spectral aperiodic exponent but not periodic alpha power were linked to white matter hyperintensities and the allostatic load index. Such index mediated the relationship between spectral aperiodic exponent and white matter hyperintensities. Thus, findings suggest that the cardiometabolic state, as measured by an allostatic load index, plays a crucial role in brain health across age, particularly influencing ventricular enlargement and increased aperiodic activity.
Article
Full-text available
Aging diminishes social cognition, and changes in this capacity can indicate brain diseases. However, the relative contribution of age, diagnosis and brain reserve to social cognition, especially among older adults and in global settings, remains unclear when considering other factors. Here, using a computational approach, we combined predictors of social cognition from a diverse sample of 1,063 older adults across nine countries. Emotion recognition, mentalizing and overall social cognition were predicted via support vector regressions from various factors, including diagnosis (subjective cognitive complaints, mild cognitive impairment, Alzheimer’s disease and behavioral variant frontotemporal dementia), demographics, cognition/executive function, brain reserve and motion artifacts from functional magnetic resonance imaging recordings. Higher cognitive/executive functions and education ranked among the top predictors, outweighing age, diagnosis and brain reserve. Network connectivity did not show predictive values. The results challenge traditional interpretations of age-related decline, patient–control differences and brain associations of social cognition, emphasizing the importance of heterogeneous factors.
Article
Full-text available
Characterization of brain states is essential for understanding its functioning in the absence of external stimuli. Brain states differ on their balance between excitation and inhibition, and on the diversity of their activity patterns. These can be respectively indexed by 1/f slope and Lempel–Ziv complexity (LZc). However, whether and how these two brain state properties relate remain elusive. Here we analyzed the relation between 1/f slope and LZc with two in-silico approaches and in both rat EEG and monkey ECoG data. We contrasted resting state with propofol anesthesia, which directly modulates the excitation-inhibition balance. We found convergent results among simulated and empirical data, showing a strong, inverse and non trivial monotonic relation between 1/f slope and complexity, consistent at both ECoG and EEG scales. We hypothesize that differentially entropic regimes could underlie the link between the excitation-inhibition balance and the vastness of the repertoire of brain systems.
Article
Full-text available
Electroencephalographic (EEG) microstates can provide a unique window into the temporal dynamics of large‐scale brain networks across brief (millisecond) timescales. Here, we analysed fundamental temporal features of microstates extracted from the broadband EEG signal in a large ( N = 139) cohort of children spanning early‐to‐middle childhood (4–12 years of age). Linear regression models were used to examine if participants' age and biological sex could predict the temporal parameters GEV , duration , coverage , and occurrence , for five microstate classes (A–E) across both eyes‐closed and eyes‐open resting‐state recordings. We further explored associations between these microstate parameters and posterior alpha power after removal of the 1/ f ‐like aperiodic signal. The microstates obtained from our neurodevelopmental EEG recordings broadly replicated the four canonical microstate classes (A to D) frequently reported in adults, with the addition of the more recently established microstate class E. Biological sex served as a significant predictor in the regression models for four of the five microstate classes (A, C, D, and E). In addition, duration and occurrence for microstate E were both found to be positively associated with age for the eyes‐open recordings, while the temporal parameters of microstates C and E both exhibited associations with alpha band spectral power. Together, these findings highlight the influence of age and sex on large‐scale functional brain networks during early‐to‐middle childhood, extending understanding of neural dynamics across this important period for brain development.
Article
Full-text available
Introduction: Accumulation and interaction of amyloid-beta (Aβ) and tau proteins during progression of Alzheimer's disease (AD) are shown to tilt neuronal circuits away from balanced excitation/inhibition (E/I). Current available techniques for noninvasive interrogation of E/I in the intact human brain, for example, magnetic resonance spectroscopy (MRS), are highly restrictive (i.e., limited spatial extent), have low temporal and spatial resolution and suffer from the limited ability to distinguish accurately between different neurotransmitters complicating its interpretation. As such, these methods alone offer an incomplete explanation of E/I. Recently, the aperiodic component of neural power spectrum, often referred to in the literature as the '1/f slope', has been described as a promising and scalable biomarker that can track disruptions in E/I potentially underlying a spectrum of clinical conditions, such as autism, schizophrenia, or epilepsy, as well as developmental E/I changes as seen in aging. Methods: Using 1/f slopes from resting-state spectral data and computational modeling, we developed a new method for inferring E/I alterations in AD. Results: We tested our method on recent freely and publicly available electroencephalography (EEG) and magnetoencephalography (MEG) datasets of patients with AD or prodromal disease and demonstrated the method's potential for uncovering regional patterns of abnormal excitatory and inhibitory parameters. Discussion: Our results provide a general framework for investigating circuit-level disorders in AD and developing therapeutic interventions that aim to restore the balance between excitation and inhibition.
Article
Full-text available
Latin American populations may present patterns of sociodemographic, ethnic and cultural diversity that can defy current universal models of healthy aging. The potential combination of risk factors that influence aging across populations in Latin American and Caribbean (LAC) countries is unknown. Compared to other regions where classical factors such as age and sex drive healthy aging, higher disparity-related factors and between-country variability could influence healthy aging in LAC countries. We investigated the combined impact of social determinants of health (SDH), lifestyle factors, cardiometabolic factors, mental health symptoms and demographics (age, sex) on healthy aging (cognition and functional ability) across LAC countries with different levels of socioeconomic development using cross-sectional and longitudinal machine learning models (n = 44,394 participants). Risk factors associated with social and health disparities, including SDH (β > 0.3), mental health (β > 0.6) and cardiometabolic risks (β > 0.22), significantly influenced healthy aging more than age and sex (with null or smaller effects: β < 0.2). These heterogeneous patterns were more pronounced in low-income to middle-income LAC countries compared to high-income LAC countries (cross-sectional comparisons), and in an upper-income to middle-income LAC country, Costa Rica, compared to China, a non-upper-income to middle-income LAC country (longitudinal comparisons). These inequity-associated and region-specific patterns inform national risk assessments of healthy aging in LAC countries and regionally tailored public health interventions.
Article
Full-text available
Bodily rhythms such as respiration are increasingly acknowledged to modulate neural oscillations underlying human action, perception, and cognition. Conversely, the link between respiration and aperiodic brain activity – a non-oscillatory reflection of excitation-inhibition (E:I) balance – has remained unstudied. Aiming to disentangle potential respiration-related dynamics of periodic and aperiodic activity, we applied recently developed algorithms of time-resolved parameter estimation to resting-state MEG and EEG data from two labs (total N = 78 participants). We provide evidence that fluctuations of aperiodic brain activity (1/f slope) are phase-locked to the respiratory cycle, which suggests that spontaneous state shifts of excitation-inhibition balance are at least partly influenced by peripheral bodily signals. Moreover, differential temporal dynamics in their coupling to non-oscillatory and oscillatory activity raise the possibility of a functional distinction in the way each component is related to respiration. Our findings highlight the role of respiration as a physiological influence on brain signalling.
Article
Full-text available
Statistical power in cognitive neuroimaging experiments is often very low. Low sample size can reduce the likelihood of detecting real effects (false negatives) and increase the risk of detecting non-existing effects by chance (false positives). Here we document our experience of leveraging a relatively unexplored method of collecting a large sample size for simple electroencephalography (EEG) studies: by recording EEG in the community during public engagement and outreach events. We collected data from 346 participants (189 females, age range 6-76 years) over 6 days, totalling 29 hours, at local science festivals. Alpha activity (6-15 Hz) was filtered from 30 seconds of signal, recorded from a single electrode placed between the occipital midline (Oz) and inion (Iz) while participants rested with their eyes closed. A total of 289 good quality datasets were obtained. Using this community-based approach, we were able to replicate controlled, lab-based findings: IAF increased during childhood, reaching a peak frequency of 10.28 Hz at 28.1 years old, and slowed again in middle and older age. Total alpha power decreased linearly, but the aperiodic-adjusted alpha power did not change over the lifespan. Aperiodic slopes and intercepts were highest in the youngest participants. There were no associations between these EEG indexes and self-reported fatigue, measured by the Multidimensional Fatigue Inventory. Finally, we present a set of important considerations for researchers who wish to collect EEG data within public engagement and outreach environments.
Article
Full-text available
Oscillatory power and phase synchronization map neuronal dynamics and are commonly studied to differentiate the healthy and diseased brain. Yet, little is known about the course and spatial variability of these features from early adulthood into old age. Leveraging magnetoencephalography (MEG) resting-state data in a cross-sectional adult sample (n = 350), we probed lifespan differences (18-88 years) in connectivity and power and interaction effects with sex. Building upon recent attempts to link brain structure and function, we tested the spatial correspondence between age effects on cortical thickness and those on functional networks. We further probed a direct structure-function relationship at the level of the study sample. We found MEG frequency-specific patterns with age and divergence between sexes in low frequencies. Connectivity and power exhibited distinct linear trajectories or turning points at midlife that might reflect different physiological processes. In the delta and beta bands, these age effects corresponded to those on cortical thickness, pointing to co-variation between the modalities across the lifespan. Structure-function coupling was frequency-dependent and observed in unimodal or multimodal regions. Altogether, we provide a comprehensive overview of the topographic functional profile of adulthood that can form a basis for neurocognitive and clinical investigations. This study further sheds new light on how the brain's structural architecture relates to fast oscillatory activity.
Article
Full-text available
Background: Cognitive abilities, particularly memory, normally decline with age. However, some individuals, often designated as superagers, can reach late life with the memory function of individuals 30 years younger. We aimed to characterise the brain structure of superagers and identify demographic, lifestyle, and clinical factors associated with this phenotype. Methods: We selected cognitively healthy participants from the Vallecas Project longitudinal cohort recruited between Oct 10, 2011, and Jan 14, 2014, aged 79·5 years or older, on the basis of their delayed verbal episodic memory score. Participants were assessed with the Free and Cued Selective Reminding Test and with three non-memory tests (the 15-item version of the Boston Naming Test, the Digit Symbol Substitution Test, and the Animal Fluency Test). Participants were classified as superagers if they scored at or above the mean values for a 50-56-year-old in the Free and Cued Selective Reminding Test and within one standard deviation of the mean or above for their age and education level in the three non-memory tests, or as typical older adults if they scored within one standard deviation of the mean for their age and education level in the Free and Cued Selective Reminding Test. Data acquired as per protocol from up to six yearly follow-ups were used for longitudinal analyses. Findings: We included 64 superagers (mean age 81·9 years; 38 [59%] women and 26 [41%] men) and 55 typical older adults (82·4 years; 35 [64%] women and 20 [36%] men). The median number of follow-up visits was 5·0 (IQR 5·0-6·0) for superagers and 5·0 (4·5-6·0) for typical older adults. Superagers exhibited higher grey matter volume cross-sectionally in the medial temporal lobe, cholinergic forebrain, and motor thalamus. Longitudinally, superagers also showed slower total grey matter atrophy, particularly within the medial temporal lobe, than did typical older adults. A machine learning classification including 89 demographic, lifestyle, and clinical predictors showed that faster movement speed (despite no group differences in exercise frequency) and better mental health were the most differentiating factors for superagers. Similar concentrations of dementia blood biomarkers in superager and typical older adult groups suggest that group differences reflect inherent superager resistance to typical age-related memory loss. Interpretation: Factors associated with dementia prevention are also relevant for resistance to age-related memory decline and brain atrophy, and the association between superageing and movement speed could provide potential novel insights into how to preserve memory function into the ninth decade. Funding: Queen Sofia Foundation, CIEN Foundation, Spanish Ministry of Science and Innovation, Alzheimer's Association, European Research Council, MAPFRE Foundation, Carl Zeiss Foundation, and the EU Comission for Horizon 2020. Translation: For the Spanish translation of the abstract see Supplementary Materials section.