Summary
The balance of excitation and inhibition is a key functional property of cortical microcircuits which changes through the lifespan. Adolescence is considered a crucial period for the maturation of excitation-inhibition balance. This has been primarily observed in animal studies, yet human in vivo evidence on adolescent maturation of the excitation-inhibition balance at the individual level is limited. Here, we developed an individualized in vivo marker of regional excitation-inhibition balance in human adolescents, estimated using large-scale simulations of biophysical network models fitted to resting-state functional magnetic resonance imaging data from two independent cross-sectional (N = 752) and longitudinal (N = 149) cohorts. We found a widespread relative increase of inhibition in association cortices paralleled by a relative age-related increase of excitation, or lack of change, in sensorimotor areas across both datasets. This developmental pattern co-aligned with multiscale markers of sensorimotor-association differentiation. The spatial pattern of excitation-inhibition development in adolescence was robust to inter-individual variability of structural connectomes and modeling configurations. Notably, we found that alternative simulation-based markers of excitation-inhibition balance show a variable sensitivity to maturational change. Taken together, our study highlights an increase of inhibition during adolescence in association areas using cross sectional and longitudinal data, and provides a robust computational framework to estimate microcircuit maturation in vivo at the individual level.
Keywords: Excitation-inhibition balance, Adolescence, Biophysical network modeling, Resting-state functional magnetic resonance imaging
Introduction
The vast repertoire of cortical functions emerges from a careful tuning of the interactions between excitatory (E) and inhibitory (I) neurons in microcircuits embedded in the structural scaffolding of the brain1. Excitation and inhibition, mainly transmitted via glutamate and γ-aminobutyric acid (GABA), respectively, are inseparable and balanced, that is, the inhibition generated in the cortical microcircuits is proportional to the local and incoming excitation2. This phenomenon has been observed during both responses to external stimuli3–5 and spontaneous cortical activity4,6. The E-I balance is proposed to be essential for central aspects of cortical functioning, including the dynamic stability of activity7, efficient coding of the information8, sharp tuning of sensory stimuli2, and generation of synchronous cortical oscillations in gamma and beta ranges2,6,9,10. Conversely, disturbed E-I balance can lead to cortical circuit dysfunctioning and is hypothesized as a key pathophysiological mechanism in various neuropsychiatric conditions such as schizophrenia, autism spectrum disorder, and epilepsy11–15.
Adolescence is a critical developmental period with substantial changes in the brain including maturation of the E-I balance16–18. During this period, several important changes occur in the architecture and function of excitatory and inhibitory neurons and synapses, which together are suggested to lead to a re-calibration of the E-I balance16. For instance, post-mortem histology of adolescent brains has shown a pruning of excitatory synapses within the prefrontal cortex, in rats19,20, non-human primates21,22 and humans23–25. In addition, post-mortem transcriptomic studies of the prefrontal cortex in animals and humans have indicated marked changes in the expression of genes involved in inhibitory neurons and GABAergic signaling, including parvalbumin9,26,27 and GABAA receptor subunits28–31. These transcriptomic changes are accompanied by the maturation of inhibitory function with stronger and shorter inhibitory postsynaptic currents, as observed in the prefrontal cortex of non-human primates9,30, overall indicating a relative increase of inhibitory synaptic transmission in this area16,32.
Currently available evidence on the in vivo maturation of the E-I balance in humans is limited, as the invasive methods used in animal studies are not feasible in humans. However, in vivo proxies of E-I balance have been proposed, relying on its putative macroscale functional consequences captured in functional imaging33,34 and electrophysiology35–37, or through biochemical quantification of glutamatergic or GABAergic neurotransmitters using magnetic resonance spectroscopy38–40. Such approaches are informative, but lack a certain level of mechanistic insight and detail that is observed with, for example, direct measurement of the excitatory and inhibitory input currents as done in animal research. Furthermore, studies on the development of E-I balance are often focused on selected areas, primarily within the prefrontal cortex, and the knowledge on the regional patterns of E-I balance maturation across the whole cerebral cortex is limited. Biophysical network modeling (BNM) of the brain is a promising computational technique that can bridge different scales of investigation at a whole-cortical level. It provides a tool to non-invasively derive mechanistic inferences about a hidden brain feature at the microscale, such as the E-I balance, based on the observed (empirical) in vivo data at the macroscale, and has provided valuable insights into brain (dys)function41–47. In this approach, dynamic spontaneous activity of brain areas are simulated using biologically realistic models which are informed by, for instance, the blood-oxygen-level-dependent (BOLD) signal measured during resting-state functional magnetic resonance imaging (rs-fMRI)48–51.
In this study, we aimed to investigate the adolescent maturation of the regional E-I balance in humans at the individual level, by applying the BNM approach on two independent cross-sectional and longitudinal neuroimaging datasets from the Philadelphia Neurodevelopmental Cohort (PNC) and the IMAGEN study52,53. We performed large-scale simulations of individualized BNMs44,54,55, in which models were informed by structural connectivity (SC) and functional imaging data of each subject. The subject-level precision of these models allowed for mapping the estimated E-I balance specifically in each individual using simulations that best represented their empirical data, and furthermore enabled studying within-subject maturation longitudinally. By doing so, we extended a previous study that used the BNM approach to study the development of the E-I balance in the PNC dataset at the level of age groups56. We demonstrated replicable effects across the two datasets indicating cross-sectional and longitudinal age-related increases of relative inhibition in the association areas and a lack of significant changes or even relative increases of excitation in the sensorimotor areas. We found this neurodevelopmental pattern of the shifts in E-I balance was aligned with the proposed sensorimotor-association axis of the cortical neurodevelopment18. Subsequently, given that the simulation results may be affected by various modeling and analytical choices57–59, or might be confounded by the variability of underlying structural connectome, as well as the noise within the simulations and parameter optimization, we extensively assessed and demonstrated the robustness of our simulation-based findings against these nuisances. Lastly, we contrasted our marker of the E-I balance to alternative previously-used BNM-based markers and highlight methodological and conceptual considerations regarding the usage and interpretation of these markers.
Results
Overview
We included 752 adolescents from the cross-sectional PNC dataset52 (409 female, mean age: 15.3±2.4 [10–19] years) and 149 participants from the longitudinal IMAGEN study53 (72 female, mean age: 14.4±0.4 years at baseline, 18.9±0.5 years at follow-up). Subject/session diffusion-weighted imaging (DWI) and rs-fMRI data was used to generate individual matrices of: i) structural connectome based on the density of white matter streamlines, ii) functional connectivity (FC) matrix as the correlation of the BOLD signals, and iii) functional connectivity dynamics (FCD) matrix as a measure of how the FC dynamically evolves through sliding windows of time during the scan, across 100 cortical areas60. Hereafter, we refer to FC and FCD matrices derived from the imaging data as empirical FC and FCD to distinguish them from simulated FC and FCD.
Next, we performed individualized BNM simulations and parameter optimizations for each subject/session to estimate their in silico regional measures of the E-I balance based on their in vivo imaging data (Figure 1). We applied the reduced Wong-Wang model61, which models each node as coupled excitatory and inhibitory neuronal pools, where the excitatory neuronal pools of different nodes are interconnected through the individual-specific SC. The model was controlled by global and regional free parameters which were fit to the empirical resting-state functional data of the target subject/session using the covariance matrix adaptation-evolution strategy (CMA-ES) optimization algorithm62–64. This involved running a maximum of 33600 simulations per subject/session using an efficient implementation of BNM simulations on graphical processing units (GPUs; https://cubnm.readthedocs.io). The model parameters included a global parameter which scales the strength of inter-regional coupling, in addition to regional parameters , and which characterize the connectivity weights between E and I neuronal pools within each node. Motivated by recent developments of this model65,66, we let and to vary across nodes, i.e., they were computed, independent of each other, through weighted combinations of six fixed biological maps that represent morphological, functional, genetic and neurochemical heterogeneity of the human cerebral cortex. These maps were obtained from independent healthy adult samples, and included average T1-weighted/T2-weighted ratio (T1w/T2w), average cortical thickness, principal gradient of FC (FC G1), principal axis of gene expression (Gene PC1), and average NMDA and GABAA/BZ receptor PET maps67–75 (Figure S1). Furthermore, in each simulation was determined based on an analytical-numerical feedback inhibition control (FIC) algorithm, which aimed to maintain the firing rate of excitatory neurons within a biologically plausible range of 3 Hz61,66. Following, from the optimal simulations of each subject/session we extracted the in silico input current to the excitatory neuron of each node , averaged across simulation time, which resulted in individual-specific maps. The values reflect in silico estimates of regional E-I balance, given results from the combination of excitatory input currents to each node (from itself, and from the excitatory neurons of the other nodes through the SC) balanced by local inhibitory currents. Therefore, increase of can be interpreted as a relative increase of excitation or decrease of inhibition within a model region.
Cross-sectional age-related variation of the E-I balance
Following, we studied cross-sectional age-related variation of the excitation-inhibition balance in adolescence, estimated based on , in the PNC dataset. The individualized optimal simulations of the PNC dataset showed a goodness-of-fit of 0.259±0.101 to the empirical data (Figure S2). Based on these simulations, we found widespread significant age-related decreases of in association areas within the frontal, parietal, and temporal lobes, in contrast to its age-related increases in visual and sensorimotor areas as well as the left posterior insula, after controlling for goodness-of-fit, sex, and in-scanner rs-fMRI motion, and adjusting for multiple comparisons at a false discovery rate (FDR) of 5% (Figure 2a). The effect sizes across these regions, calculated as Pearson correlation coefficient between age and controlled for the confounds, ranged between −0.255 to 0.156. We then assessed within-sample stability of the age effects across 100 subsamples of the data, each including half of the total sample with 376 subjects. The unthresholded age effects on across all pairs of subsamples were correlated with a mean correlation coefficient of 0.863±0.047, indicating high within-sample stability of the observed age effects (Figure 2c). Of note, assessing maturational differences of between males and females, we found no significant age-by-sex interactions after FDR correction.
Longitudinal changes of the E-I balance
To extend and assess the replicability of our findings in the cross-sectional PNC study, we next investigated the longitudinal E-I maturation in the independent IMAGEN dataset, including 149 participants assessed at the ages of 14 and 19 years. The individualized optimal simulations in the IMAGEN had a mean goodness-of-fit of 0.266±0.102 at the baseline and 0.231±0.113 at the follow-up session (Figure S3). Within these simulations, we found a significant longitudinal age-related decrease of in widespread association areas within the frontal, parietal and temporal lobe, and a significant increase in visual areas, controlled for goodness-of-fit, sex, in-scanner rs-fMRI motion, and site, and adjusted for multiple comparisons at FDR of 5% (Figure 3a). The effect sizes across these regions, calculated as the standardized mean difference of baseline to follow-up session in controlled for the confounds, ranged between −0.299 to 0.229. We next assessed the within-sample stability of age effects using 100 subsamples of the IMAGEN data, each including half of the total sample with 74 subjects. This resulted in a mean correlation of r = 0.746±0.108 between the age effect maps across all pairs of subsamples (Figure 3c). Furthermore, we found no significant age-by-sex interactions after FDR correction, yielding no evidence for sex differences in the maturation of . Of note, given the quality of the tractograms in the baseline session of the IMAGEN dataset was lower, in these simulations we used the SC of the follow-up session in the models of both sessions. However, in a subset of subjects with adequate quality of tractograms in both baseline and follow-up sessions (N = 110; 52 female), using models with session-specific SCs resulted in largely similar effects of age on (r = 0.779, pspin < 0.001; Figure S4).
We next assessed the similarity of cross-sectional and longitudinal age-related variation of observed in the two datasets. Conjunction of regions with significant age effects on in the PNC and IMAGEN datasets revealed 33 regions in the association cortices showing significant decrease of , whereas no region showed significant replicable age-related increase of (Figure 3d). Mean across these association regions, after regressing out the effects of confounds, showed a correlation coefficient of r = −0.232 with age in the PNC (T = −6.64, p < 0.001), and a standardized mean difference of −0.312 between the sessions in IMAGEN (T = −4.17, p < 0.001; Figure S5). Furthermore, the unthresholded map of longitudinal age effects in IMAGEN (Figure 3b) was significantly correlated (r = 0.630, pspin < 0.001; Figure 3e) with the cross-sectional age effect map observed in PNC (Figure 2b).
Therefore, overall, across the two datasets we observed replicable cross-sectional and longitudinal effects indicating a developmental relative increase of inhibition in the association areas in contrast to a lack of significant changes or a relative increase of excitation in sensorimotor areas.
Neurodevelopmental pattern of EI balance co-aligns with the sensorimotor-association axis of cortical organization
Having observed differential effects of age on the E-I balance across cortical areas, we next sought to investigate the embedding of this spatial neurodevelopmental pattern across different domains of cortical organization as well as meta-analytic maps of cortical function and developmental transcriptomics.
We first studied the spatial co-alignment of the observed age effects on with a previously proposed sensorimotor-association axis of cortical neurodevelopment and the multimodal cortical features it was composed of18 (Figure S6, Table S1). The age effect maps of both datasets were significantly (pspin < 0.001) correlated with the sensorimotor-association axis map (PNC: r = −0.617; IMAGEN: r = −0.607) as well as several of its components, notably including FC G1 (PNC: r = −0.691; IMAGEN: r = −0.641) and T1w/T2w (PNC: r = 0.437; IMAGEN: r = 0.548; Figure 4a, S7a). Next, comparing the age effects across seven canonical resting-state networks76, we observed more negative age effects in the default mode and frontoparietal networks compared to the somatomotor and visual networks (Figure 4b, S7b). These findings indicated co-alignment of the E-I balance maturational pattern with the sensorimotor-association axis of the cortex with higher age-related relative increase of inhibition towards the association areas.
Following, we assessed the functional relevance of the observed developmental patterns by evaluating their spatial correlation with meta-analytical maps of terms related to cognitive and behavioral functions70 obtained from the NeruoSynth database77,78 (123 terms, Table S2). We found significant (pspin < 0.001) negative correlations of the age effect maps with the maps of terms such as “intention” (PNC: r = −0.425; IMAGEN: r = −0.376) and their positive correlations with the maps of terms such as “visual perception” (PNC: r = 0.261; IMAGEN: r = 0.256), overall suggesting that areas which show a relative increase of inhibition during adolescence are associated with higher-order functions (Figure 4c, S7c).
Lastly, we performed developmental transcriptomics enrichment analysis of age effect maps. Using partial least squares regression with the gene expression maps obtained from the Allen Human Brain Atlas71,73, we identified the top 500 genes expressed higher towards the negative and positive ends of age effect maps. Next, we investigated the developmental enrichment of the two sets of genes using specific expression analysis of BrainSpan dataset79 and found that the genes expressed towards the negative ends of age effect maps to be enriched in later stages of development, primarily during and after childhood, in contrast to the genes expressed towards the positive ends of age effect maps that were enriched in earlier fetal stages of development, though not significantly (Figure 4d, S7d).
Robustness analyses
Thus far, we observed consistent age effects in the adolescent maturation of E-I balance in two independent cross-sectional and longitudinal datasets across a sensorimotor-association axis by using simulations of individualized BNMs. However, these simulation-based findings may be sensitive to various modeling and analytical choices57–59 as well as confounding effects of the underlying structural connectome or noise. Therefore, we next assessed the robustness of and its age-related changes to such nuisances, including the effects of inter-individual variability of SC, modeling configurations, and the randomness within the optimizer and simulations. For brevity and to reduce the usage of computational resources, we limited these analyses to the PNC dataset.
Inter-individual variability of structural connectome.
Using subject-specific SCs in the main analyses enabled modeling of brain function within an individualized structural scaffold which better represents each subject. However, this potentially introduces inter-individual variability of SCs as a source of variability in , particularly given is directly related to the SC in the model Eq. 1. As a result, the associations of age with model-derived features may be confounded by age-related variation of trivial features of SC, such as the node-wise strength. However, the age effect on was robust to controlling for the node-wise SC strength as an additional confound (r = 0.959, pspin < 0.001; Figure 5c, Figure S8a). Furthermore, when SC variability was eliminated by using an identical template SC in the BNM simulations of all the subjects, we observed an age effect on largely consistent with the effects observed in main analyses (r = 0.618, pspin < 0.001; Figure 5c, S8b).
Parcellation, heterogeneity maps and inter-hemispheric connections.
Following, we assessed a range of alternative modeling configurations, which we performed at the level of age groups (N = 30, each with 25–27 subjects) to limit the usage of computational resources. We found largely consistent age effect maps similar to the age effects obtained using the main analysis configuration (Figure 5b, S9) with: i) using an alternative parcellation with higher granularity of 200 nodes60 (r = 0.479, pspin < 0.001; Figure 5d, S10), ii) using two (r = 0.339, pspin < 0.001) or four (r = 0.832, pspin < 0.001) instead of six heterogeneity maps to determine regional variability of the parameters and (Figure 5e, S11), and iii) including the inter-hemispheric connections in the calculation of goodness-of-fit and the optimizer cost function (r = 0.753, pspin < 0.001; Figure 5f, S12).
Conduction velocity.
Next, we assessed the potential impact of inter-regional conduction delay at the individual level by repeating the optimal simulations while adding conduction delay between model regions informed by the subject-specific tractograms. We next calculated the node-wise median absolute deviation intraclass correlation coefficient (ICC) of between delayed-conduction and non-delayed-conduction simulations, and found its high test-retest reliability with a mean of 0.975±0.027 using velocity of 1 m/s to 0.998 ±0.003 using velocity of 6 m/s (Figure S13).
Gaussian noise random seed.
Using a similar analysis as above, we investigated the impact of the array of Gaussian noise injected into the simulations by repeating the optimal simulation of each subject with 50 alternative random seeds for the Gaussian noise. The median node-wise ICC of between the original simulations and each of the 50 alternative simulations with different seeds, had an average of 0.802±0.100 (range: [0.449–0.929]) across regions which indicates moderate to high test-retest reliability of across random seeds (Figure S14).
Optimization random seed.
It could be that within the parameter space multiple local optima exist that feature different values. To assess whether this may have been the case, we calculated the node-wise ICC of across the two optimal points obtained from the two CMA-ES runs of each subject. We found a mean ICC of 0.947±0.019 (range [0.853–0.984]) across nodes, indicating high test-retest reliability of across alternative optima, which shows a low risk of optimizers convergence to different local optima with respect to (Figure S15).
Alternative BNM-based measures of the E-I balance
In vivo estimation of the E-I balance based on BNMs has been the aim of several previous studies using this or similar models47,56,80–82. Yet, there has been no consensus on the BNM-based measures of the E-I balance and various measures have been proposed and utilized across studies. Here, we present our findings regarding alternative BNM-based measures of the E-I balance used in the previous literature and highlight some considerations regarding their usage.
Firstly, the optimal model parameters have been commonly used as measures of the E-I balance47,80–82. Variation of these parameters can be interpreted as shift of the balance towards higher excitation (e.g., increase of or ) or higher inhibition (e.g., increase of or ). However, in a multidimensional model in which these parameters can simultaneously covary and may be degenerate, the interpretation of their variations is not straight-forward. Indeed, across optimal simulations of subjects in the PNC dataset we found significant associations between optimal parameters, such as a negative association of or , indicating that lower excitatory-to-inhibitory connection weights are accompanied by (compensatory) higher inhibitory-to-excitatory connection weights (Figure 6a). These associations were also reflected in the effects of age on the parameters (Figure 6b), as for example, there was an inverse correlation between the unthresholded effects of age on or (r = −0.667, pspin < 0.001). The observed covariance between model parameters and their age effects, indicates that these age effects should not be interpreted in isolation, and based on this data, the net effect of age on the E-I balance remains ambiguous.
In our study, as a solution to this problem of degeneracy between model parameters, we focused on a state variable of model nodes within the optimal simulations, which, as a “final common pathway”, reflects the collective outcome of the various model parameters on the E-I balance within each node. At a neuronal level, the E-I balance is commonly defined as the balance, or the ratio, between excitatory and inhibitory currents, potentials or conductances onto excitatory neurons83,84. Accordingly, in our study, we quantified the E-I balance based on time-averaged input current onto the excitatory neurons, , which reflects the net difference of the excitatory and inhibitory currents onto these neurons. Of note, our measure differs from another BNM-based measure of the E-I balance based on model state variables which was utilized in a similar previous study56: the ratio of time-average excitatory synaptic gating variable, , to time-average inhibitory synaptic gating variable, . By inspecting the time series of model state variables in an example optimal simulation, we found that and are indeed positively correlated (R2 = 0.472). However, in subsequent analyses we found notable differences between these two alternative BNM-based measures of the E-I balance: (i) Assuming that the firing rate of excitatory neurons, , is an outcome of the E-I balance61 which indicates low versus high states of activity83, we expect a measure of the E-I balance to positively correlate with it. Accordingly, in an example simulation and using an exponential generalized linear mixed effects model we found that correlates positively across nodes and time with both (R2 = 0.976) and (R2 = 0.582), but is more strongly correlated with (Figure S16). This was not surprising given model Eq. 3 directly relates to . (ii) Next, given the optimal simulation of 40 randomly selected subjects of the PNC dataset, we performed perturbed simulations in which one of the model parameters was increased or decreased by 10%, pushing the simulation to an expected state of increased/decreased excitation/inhibition (e.g., 10% increase of is expected to push the system towards higher excitation). We then used paired T-tests to compare each measure of the E-I balance before and after the perturbation (Figure S17), and found larger effects of perturbations on (mean |T| = 21.476±5.375), compared to (mean |T| = 6.995±2.730). This shows that may be more sensitive to capture the variations in the E-I balance caused by parameter perturbations. (iii) Lastly, we studied the effects of age on in the PNC dataset across different model configurations and found: 1) widespread and unimodal-dominant increases using main configurations, 2) increases in unimodal and decreases in transmodal areas when using template SCs, and 3) widespread and unimodal-dominant decreases using age-group-averaged functional data, template SC, and two heterogeneity maps, which reflected the pattern observed in the study by Zhang et al.56 that used a similar configuration (Figure S18). This showed that age effects vary with different modeling configurations, though we found it was less sensitive to the effect of random seeds in the simulation (mean ICC: 0.999) and optimization (mean ICC: 0.968) relative to .
These findings, combined with the commonly used definition of the E-I balance in experimental and theoretical research at neuronal level, suggest that may be a more direct and interpretable measure of the E-I balanced using the BNM approach compared to the alternatives used in the literature. They also highlight how modeling choices and parameters can impact the outcomes of E-I changes with age.
Discussion
In this study, we used large-scale simulations of biologically realistic and individualized biophysical network models to estimate regional excitation-inhibition balance based on in vivo imaging data and evaluated its maturation during adolescence. We found a developmental shift of the E-I balance towards higher inhibition (lower excitation) in the association areas while the sensorimotor areas showed a lack of significant changes or a developmental shift towards higher excitation (lower inhibition). This finding was supported by imaging data from two independent datasets and through investigating both cross-sectional, inter-individual age-related variations of the E-I balance, as well as its longitudinal, within-individual changes through adolescence. Our observed pattern of regional variability in the E-I balance development aligned with the sensorimotor-association axis of cortical organization, and highlighted divergence of low- versus high-level brain functions, as well as early-versus late developmental timing of the sensorimotor and association areas. Lastly, we showed that our findings were robust to various modeling nuisances and choices and contrasted our simulation-based measure of the E-I balance to the alternative measures used in the literature.
The robust and replicable developmental pattern of decreased found in the association areas, indicates a relative shift of the E-I balance towards increased inhibition or decreased excitation. This observation is in line with several findings from previous animal and human studies. At the molecular level, transcriptomics and proteomics analyses have revealed peri-adolescence changes in the expression of genes related to excitation and inhibition, such as the NMDA receptor subunits85, calcium-binding proteins parvalbumin, calretinin, and calbindin, which are expressed in different types of interneurons9,26,27,32, and GABAA receptor subunits28–31. These molecular shifts mirror changes in neuronal functional properties. For instance, within the prefrontal cortex, there is an increase in the subunit composition of the GABAA receptors, from ɑ2- to ɑ1-containing receptors, which have a faster decay time, resulting in faster synaptic inhibition31,32,86. Consistent with this, recording of pyramidal neurons of the prefrontal cortex in non-human primates has indicated an increase in the strength as well as shortening of the inhibitory postsynaptic currents9,30. At the same time, microscopic investigation of the pyramidal neurons in the prefrontal cortex across different species has revealed a dramatic reduction of the excitatory synaptic density during adolescence19–23,25,87. In humans, magnetic resonance spectroscopy has been used in several studies to quantify in vivo levels of glutamate and GABA, primarily in the frontal areas. Yet, the findings of these studies have been inconsistent and include reports of age-related decrease of glutamate39,88,89 as well as increase38,40 or decrease of GABA39, increase of glutamate-GABA balance39,89, and several null findings38–40,89–91. Furthermore, a recent study on the PNC dataset quantified the in vivo E-I ratio by modeling multivariate patterns of FC and assessing their (dis)similarity to the FC of adults receiving alprazolam, a GABAergic agonist, and reported a significant developmental increase of the E-I ratio which was specific to association areas33.
In contrast to the widespread relative increase of inhibition in association regions during adolescence, we found relative increase of excitation or no significant age-related changes in the sensorimotor areas. Human cortical maturation is suggested to unfold across a sensorimotor-association axis, with a differential temporal patterning indicating earlier maturation of the sensorimotor areas in contrast to later and more protracted maturation of the association cortices18. In line with this, we found that the spatial pattern of differential E-I balance maturation across cortical areas co-aligns with the proposed sensorimotor-association axis of neurodevelopment18, and indicated that the genes preferentially expressed in association areas are more prominent in later stages of development. The sensorimotor-association neurodevelopmental variation has been observed in cortical maturation of macrostructural features92, intracortical myelination93,94, white matter connectivity95 and functional organization18,96,97, parallel to the maturation of excitation and inhibition18. Consistent with the findings in the prefrontal cortex21,22, accelerated pruning of excitatory synapses around puberty has been observed in sensorimotor areas as well87,98,99, though synaptic pruning in association regions is protracted and peaks later than in sensorimotor areas18,24. In addition, the maturation of parvalbumin inhibitory interneurons in association areas is suggested to be more prolonged18,100. Given the differences in the neurodevelopmental timing along the sensorimotor-association axis, it may be that the E-I balance matures earlier in the sensorimotor areas before adolescence, and hence, we did not find maturational increase of relative inhibition in these areas during our study age period. In line with our observation, two other studies using human in vivo markers of E-I balance reported significant increase of inhibition markers in association areas despite no significant changes in sensorimotor regions33,40. Therefore, an intriguing area for the future research will be to investigate the maturation of model-estimated E-I balance in earlier stages of development before adolescence.
Our study extends upon a recent modeling study which found widespread relative increase of inhibition across the cortex, most prominently in the sensorimotor areas, by using BNMs constructed for 29 age groups of the PNC dataset and based on a template SC of an adult sample56. In contrast, here we used large-scale simulations to construct individualized BNMs44,54,55, which allowed more specific simulation-based mapping of the E-I balance in each individual subject and this also enabled studying changes of the E-I balance longitudinally within the same individual. In addition, individualized BNMs are shown to enhance reliability of model parameters and fingerprinting accuracy of the simulated data55. Though our findings within the association areas were in agreement with the study by Zhang et al.56, indicating a maturational shift of the E-I balance towards inhibition, they diverged in the sensorimotor areas. We suspect this divergence can be attributed to several differences of the two studies, which, in addition to the usage of group-level versus individualized BNMs, include different simulation-based markers of the E-I balance as well as the methodological details of image processing, modeling and optimization. Importantly, we showed that we can (qualitatively) replicate the findings of Zhang et al.56 at the level of age groups when similarly using T1w/T2w and FC G1 as the heterogeneity maps and as the marker of E-I balance. However, we presented findings which highlighted that compared to , as well as model parameters, may be a more direct BNM-based marker of the E-I balance and be closer to its common definition in the literature as the balance, or the ratio, between excitatory and inhibitory currents, potentials or conductances onto excitatory neurons83,84. Indeed, establishing the optimal methodological choices in mapping the E-I balance using BNMs is a challenging open question, and it will be crucial for the future research to: i) comprehensively assess how modeling choices may affect the simulation-derived E-I measures, ii) evaluate the replicability of our findings in alternative datasets, particularly with higher imaging quality and more extensive follow-ups, and the fitting of BNMs to alternative or additional empirical measures, for instance using magneto-/electroencephalography44,101, and iii) ultimately, experimentally validate the E-I measures obtained through fitting of BNMs to the fMRI data against empirical measures obtained using electrical recordings, or evaluate them in response to excitatory/inhibitory interventions.
Overall, in this study we observed in vivo evidence based on individualized BNMs suggesting a replicable and robust relative increase of cortical inhibition in association areas during adolescence. The normative maturation of E-I balance is suggested to have important functional consequences18,102 and its dysmaturation is believed to be associated with various neurodevelopmental disorders11,12. For example, the neurodevelopmental model of schizophrenia suggests that aberrant cortical maturation, importantly in the development of excitatory and inhibitory functions, may contribute to the emergence of the disease later in life12,13,17,103. Therefore, beyond further methodological and validation studies discussed above, future studies will be needed to investigate the clinical relevance of adolescent maturation of E-I balance in association with the risk and diagnosis of neurodevelopmental disorders such as schizophrenia. Ultimately, investigating model-estimated E-I balance and its deviations from the normative developmental trajectory may be used as a biomarker which can present an opportunity for early detection and treatment of such disorders.
Methods
This research complies with the ethical regulations as set by The Independent Research Ethics Committee at the Medical Faculty of the Heinrich Heine University Düsseldorf (Study number 2018–317). We used previously published data from sources that have received ethics approval from their respective institutions52,53,104,105.
Participants
We studied adolescents from two population-based datasets, including the cross-sectional Philadelphia Neurodevelopmental Cohort (PNC)52,104,105 and the longitudinal IMAGEN dataset53. We selected subjects and follow-up sessions within the age range of 10 to 19 years. In the IMAGEN cohort, within this age range, phenotypic assessments were conducted at the ages of 14, 16 and 19 and imaging data was acquired at the ages of 14 and 19. Next, we excluded subjects with poor quality of raw or processed imaging data, as detailed below in the section “Image processing and quality control”. In the IMAGEN, the subjects were excluded if the imaging data had poor quality in any of the baseline or follow-up sessions, except for the diffusion weighted imaging data for which subjects were selected based on the image quality of the follow-up session. Our final sample consisted of 752 adolescents from the PNC (409 female, mean age: 15.3±2.4 years) and 149 participants from IMAGEN (72 female, baseline mean age: 14.4±0.4 years, follow-up mean age: 18.9±0.5 years). The PNC data was collected at a single center in Philadelphia, while the IMAGEN data was acquired in five different centers across Europe, in Dresden (N = 58), Paris (N = 45), Mannheim (N = 30), London (N = 14) and Dublin (N = 2). Of note, in some of the robustness analyses performed on the PNC we used aggregate data of 30 age groups, which were defined as 29 groups of 25 subjects and one group of 27 subjects, after sorting the subjects by age.
Image acquisition
T1-weighted, resting-state functional magnetic resonance imaging (rs-fMRI) and diffusion weighted imaging (DWI) data was acquired using 3T scanners from different manufacturers (PNC: Siemens, IMAGEN: Siemens, Philips, General Electric). For more details on the image acquisition parameters we refer the reader to the respective publications for each dataset (PNC: Table 1 of Ref.52, IMAGEN: Supplementary Table 5 of Ref.53). Of particular relevance to our study, the repetition and acquisition times of the rs-fMRI images were respectively 3 seconds and 6:18 minutes in the PNC, and 2.2 seconds and 6:58 minutes in the IMAGEN.
Image processing and quality control
T1-weighted structural magnetic resonance imaging
T1-weighted MRI images were processed using the recon-all command of FreeSurfer (version 7.1.1; https://surfer.nmr.mgh.harvard.edu/), which includes brain extraction, registration to standard space, tissue segmentation, and cortical surface reconstruction106,107. Quality of the FreeSurfer output was controlled based on the Euler characteristic, which represents the number of cortical surface defects before correction. We excluded outlier subjects with Euler characteristic greater than of their cohort108. The FreeSurfer output and raw T1w images were subsequently used in the pipelines of rs-fMRI and DWI processing.
Resting-state functional magnetic resonance imaging
Resting-state fMRI images were preprocessed using fMRIprep (version 22.0.0; https://fmriprep.org/en/stable/), which performed brain extraction, image registration and motion correction, estimation of confounds, and when possible, susceptibility distortion and slice time corrections109. The latter two steps were omitted in the IMAGEN data given unavailability of slice timing and variability of fieldmap formats across centers. We next further processed the output of fMRIprep, which involved the following steps: 1) applying Schaefer-100 parcellation60 by taking the average signal of all vertices within each parcel at each time point, 2) removing the first 3 volumes, 3) high-pass temporal filtering > 0.013 Hz, 4) regressing out confounds including average signal of white matter and cerebrospinal fluid voxels as well as 24 motion parameters (translation and rotation in the three directions, in addition to their squares, derivatives, and squares of derivatives)110 and 5) scrubbing motion outliers, defined based on root mean squared translation > 0.25 mm. The scrubbing was done by setting the signal in motion outlier volumes to zero while Z-scoring the rest of the volumes. This approach, compared to discarding the motion outliers, preserves the temporal structure of the BOLD signal which is important in calculating dynamic functional connectivity.
We excluded subjects/sessions with high rs-fMRI in-scanner motion defined as less than 4 minutes of scan remaining after scrubbing motion outlier volumes or a time-averaged root mean square > 0.2 mm. In addition, we performed visual quality control of the fMRIprep output and excluded subjects with gross misregistration or incomplete field of view.
Functional connectivity
Functional connectivity (FC) was calculated as the Pearson correlation of BOLD signal time series between the cortical areas. In the PNC, we additionally calculated age-group FCs by taking the mean of Z-transformed FC matrices across the subjects of each age group followed by its inverse Z-transformation.
Functional connectivity dynamics
Functional connectivity dynamics (FCD) matrix represents the temporal variability of the dynamic patterns of FC computed across sliding windows of time111. To compute the FCD matrix, we initially calculated time-resolved FC matrices of sliding windows (PNC: size 30 s and step 6 s, IMAGEN: size 30.8 s and step 4.4 s). We discarded edge windows and windows with ≥ 50% motion outliers. Subsequently, we computed FCD as the correlation between lower triangular parts of window FC patterns. The distribution of values within the FCD matrix represents the amount of recurrence of time-resolved FC patterns. In the PNC, we additionally calculated age-group FCDs by concatenating lower triangles of FCD matrices of all subjects within an age group, followed by its sorting and downsampling to reduce computational costs of model fitting.
Diffusion-weighted imaging
DWI images were processed using Micapipe (version 0.1.1; https://micapipe.readthedocs.io/en/latest/)112 which combines tools from FSL (version 6.0.0)113 and MRTrix3 (version 3.0.0)114. This pipeline performed DWI processing steps including rigid-body alignment of images, MP-PCA denoising, Gibbs ringing correction, motion and eddy current-induced distortions correction, non uniformity bias correction, registration to the processed structural image, brain mask generation and estimation of fiber orientation distributions using spherical deconvolution. Following, on each image probabilistic tractography was performed using iFOD2 algorithm to estimate 10 million streamlines115. Additionally, a track density image was computed using the iFOD1 algorithm with 1 million streamlines which was used for the quality control116. The quality control of the tractograms was done by visual inspection of the tractogram density images.
Structural connectivity
The structural connectome (SC) matrix for each subject was created using Micapipe by parcellating the tractogram using the Schaefer-100 parcellation map60 non-linearly registered to the DWI space. We subsequently normalized each SC matrix by division by its mean × 100, resulting in an equal mean of 0.01 in all SCs.
In addition to subject- and session-specific SCs of the adolescents, we used a higher-quality DWI data (3T, three shells, 140 directions) of an adult sample of 50 healthy volunteers (MICs dataset, 23 female, mean age: 29.5±5.6 years) to construct a template SC117. To do so, we obtained the SC of individual MICs subjects preprocessed using Micapipe and calculated a group-averaged SC by taking the median of streamline counts in each edge. The template SC was subsequently normalized by its mean × 100, similar to the subject- and session-specific SCs of the adolescents.
In the individualized models reported in the main analyses, we used subject-specific SCs, but the adult template SC was used in the robustness analyses on the confounding effects of inter-individual variability in SC, as well as the effects of modeling choices. Of note, in IMAGEN, given the lower quality of tractograms in the baseline session, the main analyses were performed by using the follow-up SC of each subject for the modeling of functional data in both the baseline and follow-up sessions. However, in a supplementary analysis, we additionally performed simulations using session-specific SCs, within a subset of the IMAGEN subjects with adequate quality of tractograms in both sessions (N = 110, 52 female).
Biophysical network modeling
Next, we performed BNM simulation-optimization at the level of each individual subject/session (Figure 1).
Model simulation
We simulated the spontaneous neuronal activity of 100 cortical regions from the Schaefer atlas60 as network nodes regulated by the reduced Wong-Wang model and interconnected through the SC61. In short, this model describes the activity of large ensembles of interconnected excitatory and inhibitory spiking neurons in each area by a dynamic mean field model as a reduced set of dynamic equations governing the activity of coupled excitatory (E) and inhibitory (I) pools of neurons. In this reduced model, the excitatory synaptic currents are mediated by the N-methyl-D-aspartate (NMDA) receptors and the inhibitory synaptic currents are mediated by the γ-aminobutyric acid type A (GABAA) receptors. Within each cortical region, the E and I neuronal pools are interconnected, and between regions, the E neuronal pools are coupled through a scaled SC matrix.
Model equations.
The model is mathematically described by a set of dynamic equations61. The total input current (in nA) to each excitatory () and inhibitory () neuronal pool of each cortical node , , is calculated as:
(1) |
(2) |
where and are the baseline input currents; denote the synaptic gating variables; is the number of nodes; is the structural connectivity matrix, which together with the NMDA receptor conductance, , and (global coupling), a free parameter of the model, determine the excitatory input current transmitted from the other nodes; is the recurrent excitatory connection weight; indicates the excitatory-to-inhibitory connection weight; is the inhibitory-to-excitatory connection weight; denotes recurrent inhibitory connection weight. The local connection weights , and can vary across nodes and simulations through the free parameters of the model and the feedback inhibition control, as described below.
The total input current received by each neuronal pool is subsequently transferred to , firing rates in Hz, using the sigmoidal neural response function, :
(3) |
(4) |
where and determine the slope of ; and define the thresholds above which the firing rates increase linearly with the the input currents; and determine the shape of curvature around .
Finally, the synaptic gating variables, , follow:
(5) |
(6) |
where is mediated by NMDA receptors with a decay time constant and , and is mediated by GABA receptors with a decay time constant . is uncorrelated standard Gaussian noise with an amplitude of . is subsequently bound within the range [0, 1]. Note that the fixed parameters in equations 1–6 are based on a previous paper by Deco et al.61.
Model free parameters.
The model is controlled by 15 free parameters, including , as well as bias terms and coefficients for a fixed set of six biological maps which together determine the regional values of and . More specifically, , is calculated as:
(7) |
where , the bias term, and , a vector of six coefficients, are free parameters. is a 6×100 matrix including the Z-scored biological maps in Schaefer-100 parcellation (Figure S1). These maps were based on independent samples of healthy individuals, obtained from neuromaps74 and Hansen et al.70 and included: 1) group-averaged T1-weighted/T2-weighted (T1w/T2w) ratio map of the Human Connectome Project (HCP) dataset68,69,118, 2) group-averaged cortical thickness map of the HCP dataset, 3) principal gradient of functional connectivity (FC G1)72, 4) principal axis of Allen Human Brain Atlas gene expression data (Gene PC1)71,73, 5) NMDA receptor density positron emission tomography map67,70 and 6) GABAA/BZ receptor density positron emission tomography map70,75. The resulting maps were subsequently shifted if needed to ensure . We used the following ranges for the model free parameters: : [0.5, 4.0], : [0.05, 0.75]. The range of coefficients for each map was defined as , corresponding to , , , , and .
Feedback inhibition control.
Feedback inhibition control (FIC) was used to determine the regional values of in each simulation, given the SC and other model parameters. The FIC algorithm aims to maintain a state of E-I balance in each region by adjusting to satisfy an excitatory firing rate close to 3 Hz, which is suggested to be within the biological range61. We used a two-stage implementation of the FIC by combining the original numerical implementation61 with an analytical solution proposed by Demirtaş et al.66. The latter solution analytically solves for to satisfy the self-consistency of the model equations at the steady-state condition with , corresponding to and
(8) |
in which the steady-state inhibitory synaptic gating variable was estimated by solving for in:
(9) |
Subsequently, analytical estimates of values were fed into the numerical implementation of FIC and were adjusted numerically61. In this approach, given the analytical estimates of , the model equations 1–6 are numerically integrated for a short period of 10 s and subsequently the average input current to the excitatory pool of each brain region, is calculated. If in a region exceeded the target −0.026 by more than 0.005 nA, is up-/downregulated when the input current is higher/lower than the target, and the simulation is repeated with the adjusted values in the next trial. This procedure is repeated for 10 trials or until the FIC target is satisfied in all nodes. Note that the maximum number of FIC numerical adjustment trials used here is lower than that of the original implementation to facilitate the scaling of the simulations. Furthermore, as the initial values are estimated analytically rather than being fixed to 1 (as was done in the original implementation), a smaller number of trials will be needed.
Hemodynamics model.
The simulated synaptic activity of the excitatory population of each node, , was subsequently fed to the Balloon-Windkessel model of hemodynamics to simulate BOLD signal119. This model is mathematically described by the following system of differential equations with state variables (vasodilatory signal), (blood inflow), (blood volume) and (deoxyhemoglobin content):
(10) |
(11) |
(12) |
(13) |
where is rate of signal decay, is rate of flow-dependent elimination, is hemodynamic transmit time, is Grubb’s exponent and is resting oxygen extraction fraction. These parameters were based on a previous paper by Friston et al.119. Finally, based on the model state variables the BOLD signal is calculated as: ,
(14) |
in which is the resting blood volume fraction119 and , and are dimensionless parameters that were derived for 3T scans66,120. Last, the simulated BOLD signal was downsampled to match the repetition time of the empirical rs-fMRI data, that is, 3 s for the PNC and 2.2 s for IMAGEN.
Numerical integration of the models.
For each simulation, the model equations were numerically integrated using the Euler method with a time step of 0.1 ms for the neuronal model (Equations 1–6) and a time step of 1 ms for the hemodynamic model (Equations 10–13). The model simulations were performed using in-house code (https://github.com/amnsbr/bnm_cuda, https://cubnm.readthedocs.io) on graphical processing units (GPUs) of JURECA-DC, a supercomputer at the Jülich Supercomputing Centre121. This GPU implementation enabled efficient parallelization of calculations for individual simulations (across GPU ‘blocks’) and the regions within each simulation (across GPU ‘threads’). To match the duration of empirical rs-fMRI scans, the simulations were done for a biological duration of 450 s, from which the first 30 s were discarded to ensure the BNM system’s state has stabilized.
Model evaluation
The goodness-of-fit of the simulated BOLD signal given a set of candidate parameters and SC matrix to a target empirical BOLD signal was evaluated based on three measures of static and dynamic functional connectivity, following previous studies56,65:
Static edge-level functional connectivity.
The simulated FC was calculated as the correlation of simulated BOLD signal time series between nodes. The correspondence of simulated and empirical FC patterns was evaluated by calculating the Pearson correlation coefficient between the lower triangles of the matrices (), with larger values representing higher correspondence.
Static global functional connectivity.
The absolute difference of the averaged simulated and empirical FC matrices across all the lower triangular edges () was calculated to assess the similarity of global FC strength, with smaller values showing higher correspondence.
Functional connectivity dynamics.
The simulated FCD matrix was constructed by calculating the correlation of FC patterns between sliding windows of simulated BOLD signals, as described previously for the empirical data. The correspondence of simulated and empirical FCD distributions was calculated as the Kolmogorov-Smirnov (KS) distance of their lower triangular parts (), with smaller values showing higher similarity of the distributions.
Subsequently, these measures were combined into a single measure of goodness-of-fit as . Of note, in goodness-of-fit calculations, following Demirtaş et al.66, we excluded the interhemispheric connections. However, we also performed a robustness analysis in which these connections were included in the goodness-of-fit calculations.
Parameter optimization
The model free parameters (N = 15) were fit to the empirical data of each subject/session using the covariance matrix adaptation-evolution strategy (CMA-ES) optimization algorithm62–64. CMA-ES is an efficient evolutionary optimization algorithm which features a set of Λ particles exploring the parameter space collaboratively in an iterative process. The particles from each iteration, which are individual simulations with different free parameters, are regarded as a generation from which only the best particles are selected to form the descendant population of the next generation. Specifically, at each generation the cost function of each particle is calculated, as described below. Following, a weighted mean of the best fitting ⌊Λ/2⌋ particles is calculated. Then, a new generation of particles is created by taking Λ samples from a multivariate normal distribution centered around the weighted mean of the best fitting ⌊Λ/2⌋ particles from the previous generation. The covariance is determined by a matrix which is updated to take the location of the current best points into account. In this way, the search distribution is adapted iteratively towards a concentration around the optimal solutions. This iterative process is continued for a maximum of 80 generations, following Wischnewski et al.64, and eventually, the optimal point across all generations is selected as the optimal parameters for the best fit of the simulation to the given target empirical data. We also applied an early termination rule in which the iteration was stopped if there was no improvement in the cost function greater than 0.005 over the past 30 generations.
The optimization goal was to maximize the goodness-of-fit while minimizing a penalty term. Particles were penalized if 1) the parameter of sampled particles fell outside the pre-specified ranges, in which case the parameters were corrected and a penalty was added63, or 2) the was outside the range of 2–4 Hz in any node, indicating insufficiency of the FIC. For the latter, the FIC penalty was calculated as:
(15) |
in which is the number of nodes and the summation is done across nodes with out-of-range . Of note, we refrained from setting the success of FIC as a hard constraint, but through this penalty term applied it as a soft constraint, to allow for inter-individual variability of E-I balance while keeping it within a biologically viable range61.
Given the relatively high dimensionality of the optimization problem in our model and to sufficiently cover the large parameter space, we chose Λ = 210 in the CMA-ES. Consequently, with maximum 80 generations, this involved performing a maximum of 16800 simulations per run for each subject/session, which necessitated an efficient GPU-based implementation. In addition, for each given subject/session, the model simulation-optimization was repeated twice using different random seeds of the optimization to assess and reduce the risk of local optima. We then took the best fitting simulation across the two runs of each subject/session as their optimal simulation for the next step.
Estimation of the excitation-inhibition balance in silico
Thus far, we described the procedure for deriving the optimal parameters that result in a simulation best fitting to the empirical rs-fMRI data of a given subject/session using individualized BNMs. Following, given the optimal simulation for each subject/session, we extracted an in silico measure of regional E-I balance. To do so, we calculated the average of total input current to the excitatory neurons of each region, after discarding the initial 30 seconds of the simulation. This measure can be interpreted as an in silico marker of the regional E-I balance, as (Eq. 1) results from the combination of excitatory input currents to each node (from itself, and from the excitatory neurons of the other nodes through the SC) balanced by the inhibitory currents from the inhibitory neuron of the same node. Therefore, an increase of can be interpreted as a relative increase of excitation or decrease of inhibition. Furthermore, using a similar approach we calculated as an alternative marker of the E-I balance used in a previous study56.
Perturbed simulations
We performed a control analysis to assess the effect of known perturbations in the model parameters on the alternative markers of the E-I balance, namely, and . In this analysis, we randomly selected 40 subjects of the PNC dataset and for each subject, given their optimal simulations, we performed perturbed simulations in which one of the model parameters (, , or ) was increased or decreased by 10% while the other three parameters were fixed to the optimal values. Notably, in these simulations, when was not perturbed, it was not readjusted using FIC, and was fixed to the values obtained by the FIC run on the original optimal simulation. Similarly, perturbation of was done by a 10% increase or decrease of these original values. This was to ensure only one parameter in the model is perturbed and therefore the net direction of the effect of perturbation on the E-I balance is predictable. Lastly, the effect of perturbation on the E-I balance markers in each subject was assessed using paired T-tests comparing the marker across nodes before and after the perturbation.
Statistical and contextualization analyses
Age effects
Given the regional in silico regional measures of the E-I balance for each subject/session, , we performed group-level univariate statistical analyses to investigate the effects of age on these measures. Linear regression models were used to study the effect of age on and map coefficients, with goodness-of-fit, sex, and rs-fMRI in-scanner motion (based on time-averaged root mean squared translation) as confounds. In IMAGEN, longitudinal variation of E-I measures across the two sessions was assessed using similar mixed effects linear regression models with random intercepts per subject and inclusion of site as an additional confound. In each model we excluded outliers with a dependent variable ≥3 SDs above/below the mean. We adjusted for multiple comparisons across regions using false discovery rate (FDR) based on the Benjamini/Hochberg method (q < 0.05). Similar models were used to investigate the effects of age on optimal model parameters and . We used statsmodels122 (https://www.statsmodels.org/stable/index.html) to perform regressions and FDR adjustment.
Within-sample stability of age effects using subsampling
In each dataset, we randomly selected 100 subsamples of the subjects (N within each subsample: 376 in PNC, 74 in IMAGEN) and investigated the age effects of separately in each subsample. Subsequently, we calculated the correlation of unthresholded age effect maps between all pairs of subsamples, and reported its distribution as a measure of within-sample stability.
Spatial correlation of maps
We calculated spatial correlation of maps using spin permutation in i) assessing between-sample replicability of PNC and IMAGEN age effect maps, ii) assessing stability of age effect maps within the PNC in the different robustness analyses in comparison to the age effect map observed in the main analysis, iii) assessing the correlation of age effect maps with the sensorimotor-association axis of cortical organization described in a recent study by Sydnor et al.18 as well as its components or their substitutes68,69,71–73,77,123–128 (Figure S6, Table S1), and iv) NeuroSynth77 meta-analytical maps of 123 cognitive and behavioral terms (Table S2, based on Hansen et al.70) obtained from BrainStat78. The target maps in the latter two analyses were available in fsaverage, fsLR, or MNI152 spaces and were parcellated in their respective spaces using the Schaefer-100 parcellation60. Spin permutation was implemented at parcel level using the ENIGMA Toolbox129 (https://github.com/MICA-MNI/ENIGMA) and using 1000 permutations.
Distribution of the age effects across the canonical resting-state networks
We assessed the association of the age effect maps with the map of seven canonical resting-state networks76 using one-way ANOVA with post-hoc Bonferroni-corrected independent T-tests. To control for spatial autocorrelation we assessed the statistical significance of resulting F and T statistics using null distributions generated from 1000 spun surrogates based on parcel-level spinning implemented in the ENIGMA Toolbox129.
Partial least squares regression of gene expressions and their developmental enrichment
Regional microarray expression data were obtained from 6 post-mortem brains (1 female, age = 24.0–57.0) provided by the Allen Human Brain Atlas71 (AHBA; https://human.brain-map.org). Data was processed with the abagen toolbox73 (https://abagen.readthedocs.io/en/stable/) using the Schaefer-100 atlas60. Gene expression data from the right hemisphere was excluded due to the sparsity of samples and large number of regions with no expression data. We subsequently used scikit-learn130 (https://scikit-learn.org/stable/) and performed partial least squares regression analysis to identify gene expression patterns with high spatial co-alignment with the age effect maps within the left hemisphere. After selecting the top 500 genes with highest absolute weights, we divided them into two sets of positively- and negatively- associated genes. Subsequently, using an online tool we performed specific expression analysis across developmental stages and brain structures79 (http://doughertylab.wustl.edu/csea-tool-2/). This tool performs comparison of the gene lists against developmental expression profiles from the BrainSpan Atlas of the Developing Human Brain (http://www.brainspan.org), and for each developmental stage and brain structure reports the inverse log of Fisher's exact p-values. Here, for each set of genes we reported the inverse log of p-values based on the specificity index (pSI) threshold of 0.05 within the cerebral cortex.
Test-retest reliability of across simulations
We measured node-level test-retest reliability of across simulations of the same subject performed using i) alternative optimizer seeds, ii) alternative Gaussian noise seeds, and iii) alternative conduction delays. This was done by measuring median absolute deviation intraclass correlation coefficient (ICC) of each region between the alternative simulations and across subjects (https://warwick.ac.uk/fac/sci/statistics/staff/academic-research/nichols/scripts/matlab/madicc.m).
Association of optimal model parameters across subjects
The association of optimal regional parameters (, and ) across subjects and nodes was tested via mixed effects linear regressions with random intercepts and slopes per each node. These regressions were performed using lme4 R package131. In addition, we used Pearson correlation to test the association of the optimal with the mean of optimal regional parameters.
Association of model state variables
We randomly selected one subject from the PNC dataset and evaluated the association of state variables within its optimal simulation across time points and nodes. The model state variables , and were sampled every TR after the initial 30 seconds of the simulation was removed. Subsequently, as approximates infinity when is close to zero, we excluded the data points at the top 2.5 percentile of . Subsequently we used mixed effects models with or without a logarithmic linking function to test the linear or exponential associations of with , with and with , and reported the results of the model with lower Akaike information criterion. The mixed effects models included random intercepts and slopes and were implemented using lme4 R package131 and its Python interface in pymer4 package.
Robustness analyses
We performed the following robustness analyses on the PNC dataset:
Inter-individual variability of structural connectome
We assessed the potential effect of inter-individual variability of SCs in our findings by performing the following analyses: i) We studied the age effect of which was additionally controlled for the SC strength of each node, calculated as the row-wise sum of the SC. ii) We performed BNM simulation-optimization using subject-specific functional data as the target, but with the template SC of MICs dataset determining the connectivity of model nodes, thereby eliminating the inter-individual variability of SCs as a potential source of variability in . This higher-quality DWI data from an adult sample was chosen to additionally assess the robustness of our results to potential inaccuracies of subject-level SCs derived from relatively lower quality DWI data of the adolescent datasets.
Parcellation, heterogeneity maps and inter-hemispheric connections
In these robustness analyses, we performed the BNM simulation-optimization using alternative modeling choices, including: i) the Schaefer-200 parcellation, ii) using T1w/T2w and FC G1 as the heterogeneity maps, iii) using T1w/T2w, FC G1, NMDA and GABAA/BZ as the heterogeneity maps, and iv) including inter-hemispheric connections in the goodness-of-fit and cost calculations. To reduce the computational costs, these analyses were performed at the level of age groups by using age-group functional data and the template SC determining the connectivity of model nodes. Subsequently, we compared the goodness-of-fit measures as well as the age effect maps derived from these alternative models to the output of BNM simulation-optimization done with the default configurations on the same age-group-level data.
Effects of Gaussian noise seed and inter-regional conduction delay
These analyses were performed at the level of individual subjects, by repeating the optimal simulation of each subject obtained in the main analyses with alternative configurations, including: i) using 50 different randomization seeds for generating the Gaussian noise injected into the model (Eq. 5, 6), and ii) adding conduction delay in the signal transmission between the model nodes. For the latter, delay was calculated as SC edge length obtained from the tractography of each subject, divided by a conduction velocity. For each subject we performed six alternative simulations with variable conduction velocities in the range of {1, …, 6} m/s. Of note, in simulations with conduction delay a recent history of in all nodes needs to be stored in memory, so that the input of node to node at time can be determined based on at time points ago. Accordingly, to reduce memory needed for storing this history, we performed these simulations by updating the global input to each node at intervals of 1 millisecond, instead of 0.1 milliseconds used in the main analyses. Subsequently, we calculated the goodness-of-fit measures of the alternative simulations to the empirical data of each subject and compared them with the goodness-of-fit measures of the main simulation. In addition, we calculated the ICC of between the main simulation and each of the alternative simulations. In the case of Gaussian noise seeds, we combined the goodness-of-fit as well as ICC measures of the 50 different seeds by taking their median.
Supplementary Material
Acknowledgements
AS and SLV were funded by the Max Planck Society (Otto Hahn award) and Helmholtz Association’s Initiative and Networking Fund under the Helmholtz International Lab grant agreement InterLabs-0015, and the Canada First Research Excellence Fund (CFREF Competition 2, 2015–2016) awarded to the Healthy Brains, Healthy Lives initiative at McGill University, through the Helmholtz International BigBrain Analytics and Learning Laboratory (HIBALL). LDL was supported by the Federal Ministry of Education and Research (BMBF) and the Max Planck Society (MPG), Germany. SBE was supported by the Deutsche Forschungsgemeinschaft (DFG, EI 816/21–1), the National Institute of Mental Health (R01-MH074457), and the European Union’s Horizon 2020 Research and Innovation Programme under Grant Agreement No. 945539 (HBP SGA3).
The authors gratefully acknowledge computing time on the supercomputer JURECA121 at Forschungszentrum Jülich under grant no. “eidev”. This research has been conducted using data from the dbGaP Study Accession: phs000607.v3.p2. Support for the collection of the data for Philadelphia Neurodevelopment Cohort (PNC) was provided by grant RC2MH089983 awarded to Raquel Gur and RC2MH089924 awarded to Hakon Hakonarson. Subjects were recruited and genotyped through the Center for Applied Genomics (CAG) at The Children's Hospital in Philadelphia (CHOP). Phenotypic data collection occurred at the CAG/CHOP and at the Brain Behavior Laboratory, University of Pennsylvania. This study additionally used the IMAGEN dataset, which received support from the following sources: the European Union-funded FP6 Integrated Project IMAGEN (Reinforcement-related behaviour in normal brain function and psychopathology) (LSHM-CT- 2007–037286), the Horizon 2020 funded ERC Advanced Grant ‘STRATIFY’ (Brain network based stratification of reinforcement-related disorders) (695313), Horizon Europe ‘environMENTAL’, grant no: 101057429, UK Research and Innovation (UKRI) Horizon Europe funding guarantee (10041392 and 10038599), Human Brain Project (HBP SGA 2, 785907, and HBP SGA 3, 945539), the Chinese government via the Ministry of Science and Technology (MOST). The German Center for Mental Health (DZPG), the Bundesministerium für Bildung und Forschung (BMBF grants 01GS08152; 01EV0711; Forschungsnetz AERIAL 01EE1406A, 01EE1406B; Forschungsnetz IMAC-Mind 01GL1745B), the Deutsche Forschungsgemeinschaft (DFG grants SM 80/7–2, SFB 940, TRR 265, NE 1383/14–1), the Medical Research Foundation and Medical Research Council (grants MR/R00465X/1 and MR/S020306/1), the National Institutes of Health (NIH) funded ENIGMA-grants 5U54EB020403–05, 1R56AG058854–01 and U54 EB020403 as well as NIH R01DA049238, the National Institutes of Health, Science Foundation Ireland (16/ERCD/3797). NSFC grant 82150710554. Further support was provided by grants from: - the ANR (ANR-12-SAMA-0004, AAPG2019 - GeBra), the Eranet Neuron (AF12-NEUR0008–01 - WM2NA; and ANR-18-NEUR00002–01 - ADORe), the Fondation de France (00081242), the Fondation pour la Recherche Médicale (DPA20140629802), the Mission Interministérielle de Lutte-contre-les-Drogues-et-les-Conduites-Addictives (MILDECA), the Assistance-Publique-Hôpitaux-de-Paris and INSERM (interface grant), Paris Sud University IDEX 2012, the Fondation de l’Avenir (grant AP-RM-17–013 ), the Fédération pour la Recherche sur le Cerveau.
Footnotes
Disclosures
Dr Banaschewski served in an advisory or consultancy role for eye level, Infectopharm, Medice, Neurim Pharmaceuticals, Oberberg GmbH and Takeda. He received conference support or speaker’s fee by Janssen, Medice and Takeda. He received royalties from Hogrefe, Kohlhammer, CIP Medien, Oxford University Press. The present work is unrelated to the above grants and relationships. Dr Barker has received honoraria from General Electric Healthcare for teaching on scanner programming courses. Dr Poustka served in an advisory or consultancy role for Roche and Viforpharm and received speaker’s fees from Shire. She received royalties from Hogrefe, Kohlhammer and Schattauer. The present work is unrelated to the above grants and relationships. The other authors report no biomedical financial interests or potential conflicts of interest.
Data and code availability
All the code used in this study for image processing, biophysical network modeling and the statistical analyses is available in a GitHub repository (https://github.com/amnsbr/eidev). Notably, while the main biophysical network modeling in this study were performed using https://github.com/amnsbr/bnm_cuda, we recommend and have included scripts to run the simulation-optimization using the cuBNM toolbox (https://github.com/amnsbr/cuBNM, https://cubnm.readthedocs.io), which we developed using a similar core code but with an improved interface. The adolescent datasets used in this study are restricted-access and cannot be shared openly. The neuroimaging and phenotypic data of Philadelphia Neurodevelopmental Cohort is available as part of the “Neurodevelopmental Genomics: Trajectories of Complex Phenotypes” study (dbGaP Study Accession: phs000607.v3.p2). Our access to this dataset was granted under accession number 32685. The IMAGEN dataset is similarly available to researchers by submitting a data access application to https://imagen-project.org/the-imagen-dataset/. The MICs dataset is openly available at https://osf.io/j532r/.
References
- 1.Shine J.M., Müller E.J., Munn B., Cabral J., Moran R.J., and Breakspear M. (2021). Computational models link cellular mechanisms of neuromodulation to large-scale neural dynamics. Nature Neuroscience, 1–12. 10.1038/s41593-021-00824-6. [DOI] [PubMed] [Google Scholar]
- 2.Isaacson J.S., and Scanziani M. (2011). How Inhibition Shapes Cortical Activity. Neuron 72, 231–243. 10.1016/j.neuron.2011.09.027. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 3.Anderson J.S., Carandini M., and Ferster D. (2000). Orientation tuning of input conductance, excitation, and inhibition in cat primary visual cortex. J Neurophysiol 84, 909–926. 10.1152/jn.2000.84.2.909. [DOI] [PubMed] [Google Scholar]
- 4.Okun M., and Lampl I. (2008). Instantaneous correlation of excitation and inhibition during ongoing and sensory-evoked activities. Nat Neurosci 11, 535–537. 10.1038/nn.2105. [DOI] [PubMed] [Google Scholar]
- 5.Tan A.Y.Y., Zhang L.I., Merzenich M.M., and Schreiner C.E. (2004). Tone-evoked excitatory and inhibitory synaptic conductances of primary auditory cortex neurons. J Neurophysiol 92, 630–643. 10.1152/jn.01020.2003. [DOI] [PubMed] [Google Scholar]
- 6.Atallah B.V., and Scanziani M. (2009). Instantaneous modulation of gamma oscillation frequency by balancing excitation with inhibition. Neuron 62, 566–577. 10.1016/j.neuron.2009.04.027. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 7.Wu Y.K., Miehl C., and Gjorgjieva J. (2022). Regulation of circuit organization and function through inhibitory synaptic plasticity. Trends in Neurosciences 0. 10.1016/j.tins.2022.10.006. [DOI] [PubMed] [Google Scholar]
- 8.Denève S., and Machens C.K. (2016). Efficient codes and balanced networks. Nat Neurosci 19, 375–382. 10.1038/nn.4243. [DOI] [PubMed] [Google Scholar]
- 9.Gonzalez-Burgos G., Miyamae T., Pafundo D.E., Yoshino H., Rotaru D.C., Hoftman G., Datta D., Zhang Y., Hammond M., Sampson A.R., et al. (2015). Functional Maturation of GABA Synapses During Postnatal Development of the Monkey Dorsolateral Prefrontal Cortex. Cerebral Cortex 25, 4076–4093. 10.1093/cercor/bhu122. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 10.Sohal V.S., Zhang F., Yizhar O., and Deisseroth K. (2009). Parvalbumin neurons and gamma rhythms enhance cortical circuit performance. Nature 459, 698–702. 10.1038/nature07991. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 11.Foss-Feig J.H., Adkinson B.D., Ji J.L., Yang G., Srihari V.H., McPartland J.C., Krystal J.H., Murray J.D., and Anticevic A. (2017). Searching for Cross-Diagnostic Convergence: Neural Mechanisms Governing Excitation and Inhibition Balance in Schizophrenia and Autism Spectrum Disorders. Biological Psychiatry 81, 848–861. 10.1016/j.biopsych.2017.03.005. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 12.Insel T.R. (2010). Rethinking schizophrenia. Nature 468, 187–193. 10.1038/nature09552. [DOI] [PubMed] [Google Scholar]
- 13.Rolls E.T., and Deco G. (2011). A computational neuroscience approach to schizophrenia and its onset. Neuroscience & Biobehavioral Reviews 35, 1644–1653. 10.1016/j.neubiorev.2010.09.001. [DOI] [PubMed] [Google Scholar]
- 14.Treiman D.M. (2001). GABAergic mechanisms in epilepsy. Epilepsia 42 Suppl 3, 8–12. 10.1046/j.1528-1157.2001.042suppl.3008.x. [DOI] [PubMed] [Google Scholar]
- 15.Xie K., Royer J., Rodriguez-Cruces R., Horwood L., Ngo A., Arafat T., Auer H., Sahlas E., Chen J., Zhou Y., et al. (2024). Pharmaco-resistant temporal lobe epilepsy gradually perturbs the cortex-wide excitation-inhibition balance. Preprint at bioRxiv, https://doi.org/10.1101/2024.04.22.590555 10.1101/2024.04.22.590555. [DOI] [Google Scholar]
- 16.Caballero A., Orozco A., and Tseng K.Y. (2021). Developmental regulation of excitatory-inhibitory synaptic balance in the prefrontal cortex during adolescence. Semin Cell Dev Biol 118, 60–63. 10.1016/j.semcdb.2021.02.008. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 17.Hoftman G.D., Datta D., and Lewis D.A. (2017). Layer 3 excitatory and inhibitory circuitry in the prefrontal cortex: Developmental trajectories and alterations in schizophrenia. Biological psychiatry 81, 862. 10.1016/j.biopsych.2016.05.022. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 18.Sydnor V.J., Larsen B., Bassett D.S., Alexander-Bloch A., Fair D.A., Liston C., Mackey A.P., Milham M.P., Pines A., Roalf D.R., et al. (2021). Neurodevelopment of the association cortices: Patterns, mechanisms, and implications for psychopathology. Neuron. 10.1016/j.neuron.2021.06.016. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 19.Gonzalez-Burgos G., Kroener S., Zaitsev A.V., Povysheva N.V., Krimer L.S., Barrionuevo G., and Lewis D.A. (2008). Functional maturation of excitatory synapses in layer 3 pyramidal neurons during postnatal development of the primate prefrontal cortex. Cereb Cortex 18, 626–637. 10.1093/cercor/bhm095. [DOI] [PubMed] [Google Scholar]
- 20.Koss W.A., Belden C.E., Hristov A.D., and Juraska J.M. (2014). Dendritic remodeling in the adolescent medial prefrontal cortex and the basolateral amygdala of male and female rats. Synapse 68, 61–72. 10.1002/syn.21716. [DOI] [PubMed] [Google Scholar]
- 21.Anderson S.A., Classey J.D., Condé F., Lund J.S., and Lewis D.A. (1995). Synchronous development of pyramidal neuron dendritic spines and parvalbumin-immunoreactive chandelier neuron axon terminals in layer III of monkey prefrontal cortex. Neuroscience 67, 7–22. 10.1016/0306-4522(95)00051-j. [DOI] [PubMed] [Google Scholar]
- 22.Bourgeois J.P., Goldman-Rakic P.S., and Rakic P. (1994). Synaptogenesis in the prefrontal cortex of rhesus monkeys. Cereb Cortex 4, 78–96. 10.1093/cercor/4.1.78. [DOI] [PubMed] [Google Scholar]
- 23.Huttenlocher P.R. (1979). Synaptic density in human frontal cortex - developmental changes and effects of aging. Brain Res 163, 195–205. 10.1016/0006-8993(79)90349-4. [DOI] [PubMed] [Google Scholar]
- 24.Huttenlocher P.R., and Dabholkar A.S. (1997). Regional differences in synaptogenesis in human cerebral cortex. J Comp Neurol 387, 167–178. . [DOI] [PubMed] [Google Scholar]
- 25.Petanjek Z., Judaš M., Šimić G., Rašin M.R., Uylings H.B.M., Rakic P., and Kostović I. (2011). Extraordinary neoteny of synaptic spines in the human prefrontal cortex. Proceedings of the National Academy of Sciences 108, 13281–13286. 10.1073/pnas.1105108108. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 26.Fish K.N., Hoftman G.D., Sheikh W., Kitchens M., and Lewis D.A. (2013). Parvalbumin-containing chandelier and basket cell boutons have distinctive modes of maturation in monkey prefrontal cortex. J Neurosci 33, 8352–8358. 10.1523/JNEUROSCI.0306-13.2013. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 27.Fung S.J., Webster M.J., Sivagnanasundaram S., Duncan C., Elashoff M., and Weickert C.S. (2010). Expression of interneuron markers in the dorsolateral prefrontal cortex of the developing human and in schizophrenia. Am J Psychiatry 167, 1479–1488. 10.1176/appi.ajp.2010.09060784. [DOI] [PubMed] [Google Scholar]
- 28.Datta D., Arion D., and Lewis D.A. (2015). Developmental Expression Patterns of GABAA Receptor Subunits in Layer 3 and 5 Pyramidal Cells of Monkey Prefrontal Cortex. Cereb Cortex 25, 2295–2305. 10.1093/cercor/bhu040. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 29.Fillman S.G., Duncan C.E., Webster M.J., Elashoff M., and Weickert C.S. (2010). Developmental co-regulation of the beta and gamma GABAA receptor subunits with distinct alpha subunits in the human dorsolateral prefrontal cortex. Int J Dev Neurosci 28, 513–519. 10.1016/j.ijdevneu.2010.05.004. [DOI] [PubMed] [Google Scholar]
- 30.Hashimoto T., Nguyen Q.L., Rotaru D., Keenan T., Arion D., Beneyto M., Gonzalez-Burgos G., and Lewis D.A. (2009). Protracted Developmental Trajectories of GABAA Receptor α1 and α2 Subunit Expression in Primate Prefrontal Cortex. Biol Psychiatry 65, 1015–1023. 10.1016/j.biopsych.2009.01.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 31.Hoftman G.D., and Lewis D.A. (2011). Postnatal Developmental Trajectories of Neural Circuits in the Primate Prefrontal Cortex: Identifying Sensitive Periods for Vulnerability to Schizophrenia. Schizophrenia Bulletin 37, 493. 10.1093/schbul/sbr029. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 32.Caballero A., and Tseng K.Y. (2016). GABAergic Function as a Limiting Factor for Prefrontal Maturation during Adolescence. Trends Neurosci 39, 441–448. 10.1016/j.tins.2016.04.010. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 33.Larsen B., Cui Z., Adebimpe A., Pines A., Alexander-Bloch A., Bertolero M., Calkins M.E., Gur R.E., Gur R.C., Mahadevan A.S., et al. (2022). A developmental reduction of the excitation:inhibition ratio in association cortex during adolescence. Science Advances. 10.1126/sciadv.abj8750. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 34.Trakoshis S., Martínez-Cañada P., Rocchi F., Canella C., You W., Chakrabarti B., Ruigrok A.N., Bullmore E.T., Suckling J., Markicevic M., et al. (2020). Intrinsic excitation-inhibition imbalance affects medial prefrontal cortex differently in autistic men versus women. Elife 9, e55684. 10.7554/eLife.55684. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 35.Medel V., Irani M., Crossley N., Ossandón T., and Boncompte G. (2023). Complexity and 1/f slope jointly reflect brain states. Sci Rep 13, 21700. 10.1038/s41598-023-47316-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 36.Snijders T.M., Milivojevic B., and Kemner C. (2013). Atypical excitation–inhibition balance in autism captured by the gamma response to contextual modulation. NeuroImage: Clinical 3, 65–72. 10.1016/j.nicl.2013.06.015. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 37.Uhlhaas P.J., Roux F., Rodriguez E., Rotarska-Jagiela A., and Singer W. (2010). Neural synchrony and the development of cortical networks. Trends Cogn Sci 14, 72–80. 10.1016/j.tics.2009.12.002. [DOI] [PubMed] [Google Scholar]
- 38.Ghisleni C., Bollmann S., Poil S.-S., Brandeis D., Martin E., Michels L., O’Gorman R.L., and Klaver P. (2015). Subcortical Glutamate Mediates the Reduction of Short-Range Functional Connectivity with Age in a Developmental Cohort. J. Neurosci. 35, 8433–8441. 10.1523/JNEUROSCI.4375-14.2015. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 39.Perica M.I., Calabro F.J., Larsen B., Foran W., Yushmanov V.E., Hetherington H., Tervo-Clemmens B., Moon C.-H., and Luna B. (2022). Development of frontal GABA and glutamate supports excitation/inhibition balance from adolescence into adulthood. Progress in Neurobiology, 102370. 10.1016/j.pneurobio.2022.102370. [DOI] [PubMed] [Google Scholar]
- 40.Silveri M.M., Sneider J.T., Crowley D.J., Covell M.J., Acharya D., Rosso I.M., and Jensen J.E. (2013). Frontal lobe γ-aminobutyric acid levels during adolescence: associations with impulsivity and response inhibition. Biol Psychiatry 74, 296–304. 10.1016/j.biopsych.2013.01.033. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 41.Cabral J., Fernandes H.M., Van Hartevelt T.J., James A.C., Kringelbach M.L., and Deco G. (2013). Structural connectivity in schizophrenia and its impact on the dynamics of spontaneous functional networks. Chaos 23, 046111. 10.1063/1.4851117. [DOI] [PubMed] [Google Scholar]
- 42.Deco G., Cruzat J., Cabral J., Knudsen G.M., Carhart-Harris R.L., Whybrow P.C., Logothetis N.K., and Kringelbach M.L. (2018). Whole-Brain Multimodal Neuroimaging Model Using Serotonin Receptor Maps Explains Non-linear Functional Effects of LSD. Curr Biol 28, 3065–3074.e6. 10.1016/j.cub.2018.07.083. [DOI] [PubMed] [Google Scholar]
- 43.Froudist-Walsh S., Bliss D.P., Ding X., Rapan L., Niu M., Knoblauch K., Zilles K., Kennedy H., Palomero-Gallagher N., and Wang X.-J. (2021). A dopamine gradient controls access to distributed working memory in the large-scale monkey cortex. Neuron 109, 3500–3520.e13. 10.1016/j.neuron.2021.08.024. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 44.Jirsa V.K., Proix T., Perdikis D., Woodman M.M., Wang H., Gonzalez-Martinez J., Bernard C., Bénar C., Guye M., Chauvel P., et al. (2017). The Virtual Epileptic Patient: Individualized whole-brain models of epilepsy spread. NeuroImage 145, 377–388. 10.1016/j.neuroimage.2016.04.049. [DOI] [PubMed] [Google Scholar]
- 45.Jung K., Florin E., Patil K.R., Caspers J., Rubbert C., Eickhoff S.B., and Popovych O.V. (2023). Whole-brain dynamical modelling for classification of Parkinson’s disease. Brain Communications 5, fcac331. 10.1093/braincomms/fcac331. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 46.Kringelbach M.L., Cruzat J., Cabral J., Knudsen G.M., Carhart-Harris R., Whybrow P.C., Logothetis N.K., and Deco G. (2020). Dynamic coupling of whole-brain neuronal and neurotransmitter systems. Proceedings of the National Academy of Sciences 117, 9566–9576. 10.1073/pnas.1921475117. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 47.Park B., Hong S.-J., Valk S.L., Paquola C., Benkarim O., Bethlehem R.A.I., Di Martino A., Milham M.P., Gozzi A., Yeo B.T.T., et al. (2021). Differences in subcortico-cortical interactions identified from connectome and microcircuit models in autism. Nat Commun 12, 2225. 10.1038/s41467-021-21732-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 48.Breakspear M. (2017). Dynamic models of large-scale brain activity. Nature Neuroscience 20, 340–352. 10.1038/nn.4497. [DOI] [PubMed] [Google Scholar]
- 49.Deco G., and Kringelbach M.L. (2014). Great Expectations: Using Whole-Brain Computational Connectomics for Understanding Neuropsychiatric Disorders. Neuron 84, 892–905. 10.1016/j.neuron.2014.08.034. [DOI] [PubMed] [Google Scholar]
- 50.Popovych O.V., Manos T., Hoffstaedter F., and Eickhoff S.B. (2019). What Can Computational Models Contribute to Neuroimaging Data Analytics? Frontiers in Systems Neuroscience 12, 68. 10.3389/fnsys.2018.00068. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 51.Ritter P., Schirner M., McIntosh A.R., and Jirsa V.K. (2013). The Virtual Brain Integrates Computational Modeling and Multimodal Neuroimaging. Brain Connectivity 3, 121–145. 10.1089/brain.2012.0120. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 52.Satterthwaite T.D., Elliott M.A., Ruparel K., Loughead J., Prabhakaran K., Calkins M.E., Hopson R., Jackson C., Keefe J., Riley M., et al. (2014). Neuroimaging of the Philadelphia Neurodevelopmental Cohort. Neuroimage 86, 544–553. 10.1016/j.neuroimage.2013.07.064. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 53.Schumann G., Loth E., Banaschewski T., Barbot A., Barker G., Büchel C., Conrod P.J., Dalley J.W., Flor H., Gallinat J., et al. (2010). The IMAGEN study: reinforcement-related behaviour in normal brain function and psychopathology. Mol Psychiatry 15, 1128–1139. 10.1038/mp.2010.4. [DOI] [PubMed] [Google Scholar]
- 54.Bansal K., Nakuci J., and Muldoon S.F. (2018). Personalized brain network models for assessing structure–function relationships. Current Opinion in Neurobiology 52, 42–47. 10.1016/j.conb.2018.04.014. [DOI] [PubMed] [Google Scholar]
- 55.Domhof J.W.M., Eickhoff S.B., and Popovych O.V. (2022). Reliability and subject specificity of personalized whole-brain dynamical models. NeuroImage 257, 119321. 10.1016/j.neuroimage.2022.119321. [DOI] [PubMed] [Google Scholar]
- 56.Zhang S., Larsen B., Sydnor V.J., Zeng T., An L., Yan X., Kong R., Kong X., Gur R.C., Gur R.E., et al. (2024). In vivo whole-cortex marker of excitation-inhibition ratio indexes cortical maturation and cognitive ability in youth. Proceedings of the National Academy of Sciences 121, e2318641121. 10.1073/pnas.2318641121. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 57.Domhof J.W.M., Jung K., Eickhoff S.B., and Popovych O.V. (2021). Parcellation-induced variation of empirical and simulated brain connectomes at group and subject levels. Network Neuroscience 5, 798–830. 10.1162/netn_a_00202. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 58.Jung K., Eickhoff S.B., and Popovych O.V. (2021). Tractography density affects whole-brain structural architecture and resting-state dynamical modeling. NeuroImage 237, 118176. 10.1016/j.neuroimage.2021.118176. [DOI] [PubMed] [Google Scholar]
- 59.Popovych O.V., Jung K., Manos T., Diaz-Pier S., Hoffstaedter F., Schreiber J., Yeo B.T.T., and Eickhoff S.B. (2021). Inter-subject and inter-parcellation variability of resting-state whole-brain dynamical modeling. NeuroImage 236, 118201. 10.1016/j.neuroimage.2021.118201. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 60.Schaefer A., Kong R., Gordon E.M., Laumann T.O., Zuo X.-N., Holmes A.J., Eickhoff S.B., and Yeo B.T.T. (2018). Local-Global Parcellation of the Human Cerebral Cortex from Intrinsic Functional Connectivity MRI. Cereb Cortex 28, 3095–3114. 10.1093/cercor/bhx179. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 61.Deco G., Ponce-Alvarez A., Hagmann P., Romani G.L., Mantini D., and Corbetta M. (2014). How local excitation-inhibition ratio impacts the whole brain dynamics. J Neurosci 34, 7886–7898. 10.1523/JNEUROSCI.5068-13.2014. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 62.Hansen N. (2009). Benchmarking a BI-population CMA-ES on the BBOB-2009 function testbed. In Proceedings of the 11th Annual Conference Companion on Genetic and Evolutionary Computation Conference: Late Breaking Papers (ACM), pp. 2389–2396. 10.1145/1570256.1570333. [DOI] [Google Scholar]
- 63.Hansen N. (2016). The CMA Evolution Strategy: A Tutorial. Preprint at arXiv, 10.48550/arXiv.1604.00772 https://doi.org/10.48550/arXiv.1604.00772. [DOI] [Google Scholar]
- 64.Wischnewski K.J., Eickhoff S.B., Jirsa V.K., and Popovych O.V. (2022). Towards an efficient validation of dynamical whole-brain models. Sci Rep 12, 4331. 10.1038/s41598-022-07860-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 65.Deco G., Kringelbach M.L., Arnatkeviciute A., Oldham S., Sabaroedin K., Rogasch N.C., Aquino K.M., and Fornito A. (2021). Dynamical consequences of regional heterogeneity in the brain’s transcriptional landscape. Science Advances 7, eabf4752. 10.1126/sciadv.abf4752. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 66.Demirtaş M., Burt J.B., Helmer M., Ji J.L., Adkinson B.D., Glasser M.F., Van Essen D.C., Sotiropoulos S.N., Anticevic A., and Murray J.D. (2019). Hierarchical Heterogeneity across Human Cortex Shapes Large-Scale Neural Dynamics. Neuron 101, 1181–1194.e13. 10.1016/j.neuron.2019.01.017. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 67.Galovic M., Erlandsson K., Fryer T.D., Hong Y.T., Manavaki R., Sari H., Chetcuti S., Thomas B.A., Fisher M., Sephton S., et al. (2021). Validation of a combined image derived input function and venous sampling approach for the quantification of [18F]GE-179 PET binding in the brain. Neuroimage 237, 118194. 10.1016/j.neuroimage.2021.118194. [DOI] [PubMed] [Google Scholar]
- 68.Glasser M.F., Sotiropoulos S.N., Wilson J.A., Coalson T.S., Fischl B., Andersson J.L., Xu J., Jbabdi S., Webster M., Polimeni J.R., et al. (2013). The minimal preprocessing pipelines for the Human Connectome Project. Neuroimage 80, 105–124. 10.1016/j.neuroimage.2013.04.127. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 69.Glasser M.F., and Van Essen D.C. (2011). Mapping human cortical areas in vivo based on myelin content as revealed by T1- and T2-weighted MRI. J Neurosci 31, 11597–11616. 10.1523/JNEUROSCI.2180-11.2011. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 70.Hansen J.Y., Shafiei G., Markello R.D., Smart K., Cox S.M.L., Nørgaard M., Beliveau V., Wu Y., Gallezot J.-D., Aumont É., et al. (2022). Mapping neurotransmitter systems to the structural and functional organization of the human neocortex. Nat Neurosci 25, 1569–1581. 10.1038/s41593-022-01186-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 71.Hawrylycz M.J., Lein E.S., Guillozet-Bongaarts A.L., Shen E.H., Ng L., Miller J.A., van de Lagemaat L.N., Smith K.A., Ebbert A., Riley Z.L., et al. (2012). An anatomically comprehensive atlas of the adult human brain transcriptome. Nature 489, 391–399. 10.1038/nature11405. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 72.Margulies D.S., Ghosh S.S., Goulas A., Falkiewicz M., Huntenburg J.M., Langs G., Bezgin G., Eickhoff S.B., Castellanos F.X., Petrides M., et al. (2016). Situating the default-mode network along a principal gradient of macroscale cortical organization. PNAS 113, 12574–12579. 10.1073/pnas.1608282113. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 73.Markello R.D., Arnatkeviciute A., Poline J.-B., Fulcher B.D., Fornito A., and Misic B. (2021). Standardizing workflows in imaging transcriptomics with the abagen toolbox. eLife 10, e72129. 10.7554/eLife.72129. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 74.Markello R.D., Hansen J.Y., Liu Z.-Q., Bazinet V., Shafiei G., Suárez L.E., Blostein N., Seidlitz J., Baillet S., Satterthwaite T.D., et al. (2022). neuromaps: structural and functional interpretation of brain maps. Nat Methods 19, 1472–1479. 10.1038/s41592-022-01625-w. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 75.Nørgaard M., Beliveau V., Ganz M., Svarer C., Pinborg L.H., Keller S.H., Jensen P.S., Greve D.N., and Knudsen G.M. (2021). A high-resolution in vivo atlas of the human brain’s benzodiazepine binding site of GABAA receptors. NeuroImage 232, 117878. 10.1016/j.neuroimage.2021.117878. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 76.Yeo B.T.T., Krienen F.M., Sepulcre J., Sabuncu M.R., Lashkari D., Hollinshead M., Roffman J.L., Smoller J.W., Zöllei L., Polimeni J.R., et al. (2011). The organization of the human cerebral cortex estimated by intrinsic functional connectivity. J Neurophysiol 106, 1125–1165. 10.1152/jn.00338.2011. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 77.Yarkoni T., Poldrack R.A., Nichols T.E., Van Essen D.C., and Wager T.D. (2011). Large-scale automated synthesis of human functional neuroimaging data. Nat Methods 8, 665–670. 10.1038/nmeth.1635. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 78.Larivière S., Bayrak Ş., Vos de Wael R., Benkarim O., Herholz P., Rodriguez-Cruces R., Paquola C., Hong S.-J., Misic B., Evans A.C., et al. (2023). BrainStat: A toolbox for brain-wide statistics and multimodal feature associations. Neuroimage 266, 119807. 10.1016/j.neuroimage.2022.119807. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 79.Dougherty J.D., Schmidt E.F., Nakajima M., and Heintz N. (2010). Analytical approaches to RNA profiling data for the identification of genes enriched in specific cells. Nucleic Acids Res 38, 4218–4230. 10.1093/nar/gkq130. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 80.Klein P.C., Ettinger U., Schirner M., Ritter P., Rujescu D., Falkai P., Koutsouleris N., Kambeitz-Ilankovic L., and Kambeitz J. (2021). Brain Network Simulations Indicate Effects of Neuregulin-1 Genotype on Excitation-Inhibition Balance in Cortical Dynamics. Cerebral Cortex 31, 2013–2025. 10.1093/cercor/bhaa339. [DOI] [PubMed] [Google Scholar]
- 81.Monteverdi A., Palesi F., Costa A., Vitali P., Pichiecchio A., Cotta Ramusino M., Bernini S., Jirsa V., Gandini Wheeler-Kingshott C.A.M., and D’Angelo E. (2022). Subject-specific features of excitation/inhibition profiles in neurodegenerative diseases. Frontiers in Aging Neuroscience 14. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 82.Yang G.J., Murray J.D., Wang X.-J., Glahn D.C., Pearlson G.D., Repovs G., Krystal J.H., and Anticevic A. (2016). Functional hierarchy underlies preferential connectivity disturbances in schizophrenia. Proc Natl Acad Sci U S A 113, E219–228. 10.1073/pnas.1508436113. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 83.Haider B., Duque A., Hasenstaub A.R., and McCormick D.A. (2006). Neocortical Network Activity In Vivo Is Generated through a Dynamic Balance of Excitation and Inhibition. J. Neurosci. 26, 4535–4545. 10.1523/JNEUROSCI.5297-05.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 84.He H., and Cline H.T. (2019). What Is Excitation/Inhibition and How Is It Regulated? A Case of the Elephant and the Wisemen. Journal of Experimental Neuroscience 13. 10.1177/1179069519859371. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 85.Henson M.A., Roberts A.C., Salimi K., Vadlamudi S., Hamer R.M., Gilmore J.H., Jarskog L.F., and Philpot B.D. (2008). Developmental regulation of the NMDA receptor subunits, NR3A and NR1, in human prefrontal cortex. Cereb Cortex 18, 2560–2573. 10.1093/cercor/bhn017. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 86.Duncan C.E., Webster M.J., Rothmond D.A., Bahn S., Elashoff M., and Shannon Weickert C. (2010). Prefrontal GABAA receptor α-subunit expression in normal postnatal human development and schizophrenia. Journal of Psychiatric Research 44, 673–681. 10.1016/j.jpsychires.2009.12.007. [DOI] [PubMed] [Google Scholar]
- 87.Bourgeois J.P., and Rakic P. (1993). Changes of synaptic density in the primary visual cortex of the macaque monkey from fetal to adult stage. J Neurosci 13, 2801–2820. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 88.Shimizu M., Suzuki Y., Yamada K., Ueki S., Watanabe M., Igarashi H., and Nakada T. (2017). Maturational decrease of glutamate in the human cerebral cortex from childhood to young adulthood: a 1H-MR spectroscopy study. Pediatr Res 82, 749–752. 10.1038/pr.2017.101. [DOI] [PubMed] [Google Scholar]
- 89.McKeon S.D., Perica M.I., Parr A.C., Calabro F.J., Foran W., Hetherington H., Moon C.-H., and Luna B. (2024). Aperiodic EEG and 7T MRSI evidence for maturation of E/I balance supporting the development of working memory through adolescence. Dev Cogn Neurosci 66, 101373. 10.1016/j.dcn.2024.101373. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 90.Blüml S., Wisnowski J.L., Nelson M.D. Jr, , Paquette L., Gilles F.H., Kinney H.C., and Panigrahy A. (2013). Metabolic Maturation of the Human Brain From Birth Through Adolescence: Insights From In Vivo Magnetic Resonance Spectroscopy. Cerebral Cortex 23, 2944–2955. 10.1093/cercor/bhs283. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 91.Gleich T., Lorenz R.C., Pöhland L., Raufelder D., Deserno L., Beck A., Heinz A., Kühn S., and Gallinat J. (2015). Frontal glutamate and reward processing in adolescence and adulthood. Brain Struct Funct 220, 3087–3099. 10.1007/s00429-014-0844-3. [DOI] [PubMed] [Google Scholar]
- 92.Bethlehem R.a. I , Seidlitz J., White S.R., Vogel J.W., Anderson K.M., Adamson C., Adler S., Alexopoulos G.S., Anagnostou E., Areces-Gonzalez A, et al. (2022). Brain charts for the human lifespan. Nature, 1–11. 10.1038/s41586-022-04554-y. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 93.Grydeland H., Vértes P.E., Váša F., Romero-Garcia R., Whitaker K., Alexander-Bloch A.F., Bjørnerud A., Patel A.X., Sederevičius D., Tamnes C.K., et al. (2019). Waves of Maturation and Senescence in Micro-structural MRI Markers of Human Cortical Myelination over the Lifespan. Cerebral Cortex 29, 1369–1381. 10.1093/cercor/bhy330. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 94.Paquola C., Bethlehem R.A., Seidlitz J., Wagstyl K., Romero-Garcia R., Whitaker K.J., Vos De Wael R., Williams G.B., Vértes P.E., Margulies D.S., et al. (2019). Shifts in myeloarchitecture characterise adolescent development of cortical gradients. eLife 8. 10.7554/eLife.50482. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 95.Simmonds D.J., Hallquist M.N., Asato M., and Luna B. (2014). Developmental stages and sex differences of white matter and behavioral development through adolescence: a longitudinal diffusion tensor imaging (DTI) study. Neuroimage 92, 356–368. 10.1016/j.neuroimage.2013.12.044. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 96.Bassett D.S., Xia C.H., and Satterthwaite T.D. (2018). Understanding the Emergence of Neuropsychiatric Disorders With Network Neuroscience. Biological Psychiatry: Cognitive Neuroscience and Neuroimaging 3, 742–753. 10.1016/j.bpsc.2018.03.015. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 97.Dong H.-M., Margulies D.S., Zuo X.-N., and Holmes A.J. (2021). Shifting gradients of macroscale cortical organization mark the transition from childhood to adolescence. Proc Natl Acad Sci U S A 118, e2024448118. 10.1073/pnas.2024448118. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 98.Zecevic N., Bourgeois J.P., and Rakic P. (1989). Changes in synaptic density in motor cortex of rhesus monkey during fetal and postnatal life. Brain Res Dev Brain Res 50, 11–32. 10.1016/0165-3806(89)90124-7. [DOI] [PubMed] [Google Scholar]
- 99.Zecevic N., and Rakic P. (1991). Synaptogenesis in monkey somatosensory cortex. Cereb Cortex 1, 510–523. 10.1093/cercor/1.6.510. [DOI] [PubMed] [Google Scholar]
- 100.Reh R.K., Dias B.G., Nelson C.A., Kaufer D., Werker J.F., Kolb B., Levine J.D., and Hensch T.K. (2020). Critical period regulation across multiple timescales. Proc Natl Acad Sci U S A 117, 23242–23251. 10.1073/pnas.1820836117. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 101.Schirner M., McIntosh A.R., Jirsa V., Deco G., and Ritter P. (2018). Inferring multi-scale neural mechanisms with brain network modelling. eLife 7, e28927. 10.7554/eLife.28927. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 102.Larsen B., and Luna B. (2018). Adolescence as a neurobiological critical period for the development of higher-order cognition. Neuroscience & Biobehavioral Reviews 94, 179–195. 10.1016/j.neubiorev.2018.09.005. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 103.Keshavan M.S., Giedd J., Lau J.Y.F., Lewis D.A., and Paus T. (2014). Changes in the adolescent brain and the pathophysiology of psychotic disorders. Lancet Psychiatry 1, 549–558. 10.1016/S2215-0366(14)00081-9. [DOI] [PubMed] [Google Scholar]
- 104.Calkins M.E., Merikangas K.R., Moore T.M., Burstein M., Behr M.A., Satterthwaite T.D., Ruparel K., Wolf D.H., Roalf D.R., Mentch F.D., et al. (2015). The Philadelphia Neurodevelopmental Cohort: constructing a deep phenotyping collaborative. J Child Psychol Psychiatry 56, 1356–1369. 10.1111/jcpp.12416. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 105.Calkins M.E., Moore T.M., Merikangas K.R., Burstein M., Satterthwaite T.D., Bilker W.B., Ruparel K., Chiavacci R., Wolf D.H., Mentch F., et al. (2014). The psychosis spectrum in a young U.S. community sample: findings from the Philadelphia Neurodevelopmental Cohort. World Psychiatry 13, 296–305. 10.1002/wps.20152. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 106.Dale A.M., Fischl B., and Sereno M.I. (1999). Cortical surface-based analysis. I. Segmentation and surface reconstruction. Neuroimage 9, 179–194. 10.1006/nimg.1998.0395. [DOI] [PubMed] [Google Scholar]
- 107.Fischl B., Sereno M.I., and Dale A.M. (1999). Cortical surface-based analysis. II: Inflation, flattening, and a surface-based coordinate system. Neuroimage 9, 195–207. 10.1006/nimg.1998.0396. [DOI] [PubMed] [Google Scholar]
- 108.Monereo-Sánchez J., de Jong J.J.A., Drenthen G.S., Beran M., Backes W.H., Stehouwer C.D.A., Schram M.T., Linden D.E.J., and Jansen J.F.A. (2021). Quality control strategies for brain MRI segmentation and parcellation: Practical approaches and recommendations - insights from the Maastricht study. Neuroimage 237, 118174. 10.1016/j.neuroimage.2021.118174. [DOI] [PubMed] [Google Scholar]
- 109.Esteban O., Markiewicz C.J., Blair R.W., Moodie C.A., Isik A.I., Erramuzpe A., Kent J.D., Goncalves M., DuPre E., Snyder M., et al. (2018). fMRIPrep: a robust preprocessing pipeline for functional MRI. Nat Methods 16, 111–116. 10.1038/s41592-018-0235-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 110.Friston K.J., Williams S., Howard R., Frackowiak R.S., and Turner R. (1996). Movement-related effects in fMRI time-series. Magn Reson Med 35, 346–355. 10.1002/mrm.1910350312. [DOI] [PubMed] [Google Scholar]
- 111.Hansen E.C.A., Battaglia D., Spiegler A., Deco G., and Jirsa V.K. (2015). Functional connectivity dynamics: Modeling the switching behavior of the resting state. NeuroImage 105, 525–535. 10.1016/j.neuroimage.2014.11.001. [DOI] [PubMed] [Google Scholar]
- 112.Cruces R.R., Royer J., Herholz P., Larivière S., Vos de Wael R., Paquola C., Benkarim O., Park B., Degré-Pelletier J., Nelson M.C., et al. (2022). Micapipe: A pipeline for multimodal neuroimaging and connectome analysis. NeuroImage 263, 119612. 10.1016/j.neuroimage.2022.119612. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 113.Jenkinson M., Beckmann C.F., Behrens T.E.J., Woolrich M.W., and Smith S.M. (2012). FSL. NeuroImage 62, 782–790. 10.1016/j.neuroimage.2011.09.015. [DOI] [PubMed] [Google Scholar]
- 114.Tournier J.-D., Smith R., Raffelt D., Tabbara R., Dhollander T., Pietsch M., Christiaens D., Jeurissen B., Yeh C.-H., and Connelly A. (2019). MRtrix3: A fast, flexible and open software framework for medical image processing and visualisation. Neuroimage 202, 116137. 10.1016/j.neuroimage.2019.116137. [DOI] [PubMed] [Google Scholar]
- 115.Tournier J.D., Calamante F., and Connelly A. (2010). Improved probabilistic streamlines tractography by 2nd order integration over fibre orientation distributions. In Proceedings of the international society for magnetic resonance in medicine (John Wiley & Sons, Inc, New Jersey; ). [Google Scholar]
- 116.Tournier J.-D., Calamante F., and Connelly A. (2012). MRtrix: Diffusion tractography in crossing fiber regions. Int. J. Imaging Syst. Technol. 22, 53–66. 10.1002/ima.22005. [DOI] [Google Scholar]
- 117.Royer J., Rodríguez-Cruces R., Tavakol S., Larivière S., Herholz P., Li Q., Vos de Wael R., Paquola C., Benkarim O., Park B., et al. (2022). An Open MRI Dataset For Multiscale Neuroscience. Sci Data 9, 569. 10.1038/s41597-022-01682-y. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 118.Van Essen D.C., Smith S.M., Barch D.M., Behrens T.E.J., Yacoub E., Ugurbil K., and WU-Minn HCP Consortium (2013). The WU-Minn Human Connectome Project: an overview. Neuroimage 80, 62–79. 10.1016/j.neuroimage.2013.05.041. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 119.Friston K.J., Harrison L., and Penny W. (2003). Dynamic causal modelling. Neuroimage 19, 1273–1302. 10.1016/s1053-8119(03)00202-7. [DOI] [PubMed] [Google Scholar]
- 120.Heinzle J., Koopmans P.J., den Ouden H.E.M., Raman S., and Stephan K.E. (2016). A hemodynamic model for layered BOLD signals. NeuroImage 125, 556–570. 10.1016/j.neuroimage.2015.10.025. [DOI] [PubMed] [Google Scholar]
- 121.Jülich Supercomputing Centre (2021). JURECA: Data Centric and Booster Modules implementing the Modular Supercomputing Architecture at Jülich Supercomputing Centre. Journal of large-scale research facilities 7. 10.17815/jlsrf-7-182. [DOI] [Google Scholar]
- 122.Seabold S., and Perktold J. (2010). Statsmodels: Econometric and statistical modeling with python. In Proceedings of the 9th Python in Science Conference (Austin, TX), pp. 10–25080. [Google Scholar]
- 123.Hill J., Inder T., Neil J., Dierker D., Harwell J., and Van Essen D. (2010). Similar patterns of cortical expansion during human development and evolution. Proc Natl Acad Sci U S A 107, 13135–13140. 10.1073/pnas.1001229107. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 124.Reardon P.K., Seidlitz J., Vandekar S., Liu S., Patel R., Park M.T.M., Alexander-Bloch A., Clasen L.S., Blumenthal J.D., Lalonde F.M., et al. (2018). Normative brain size variation and brain shape diversity in humans. Science 360, 1222–1227. 10.1126/science.aar2578. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 125.Vaishnavi S.N., Vlassenko A.G., Rundle M.M., Snyder A.Z., Mintun M.A., and Raichle M.E. (2010). Regional aerobic glycolysis in the human brain. Proc Natl Acad Sci U S A 107, 17757–17762. 10.1073/pnas.1010459107. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 126.Satterthwaite T.D., Shinohara R.T., Wolf D.H., Hopson R.D., Elliott M.A., Vandekar S.N., Ruparel K., Calkins M.E., Roalf D.R., Gennatas E.D., et al. (2014). Impact of puberty on the evolution of cerebral perfusion during adolescence. Proc Natl Acad Sci U S A 111, 8643–8648. 10.1073/pnas.1400178111. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 127.Poldrack R.A., Kittur A., Kalar D., Miller E., Seppa C., Gil Y., Parker D.S., Sabb F.W., and Bilder R.M. (2011). The cognitive atlas: toward a knowledge foundation for cognitive neuroscience. Front Neuroinform 5, 17. 10.3389/fninf.2011.00017. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 128.Saberi A., Paquola C., Wagstyl K., Hettwer M.D., Bernhardt B.C., Eickhoff S.B., and Valk S.L. (2023). The regional variation of laminar thickness in the human isocortex is related to cortical hierarchy and interregional connectivity. PLOS Biology 21, e3002365. 10.1371/journal.pbio.3002365. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 129.Larivière S., Paquola C., Park B., Royer J., Wang Y., Benkarim O., Vos de Wael R., Valk S.L., Thomopoulos S.I., Kirschner M., et al. (2021). The ENIGMA Toolbox: multiscale neural contextualization of multisite neuroimaging datasets. Nat Methods 18, 698–700. 10.1038/s41592-021-01186-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 130.Pedregosa F., Varoquaux G., Gramfort A., Michel V., Thirion B., Grisel O., Blondel M., Prettenhofer P., Weiss R., and Dubourg V. (2011). Scikit-learn: Machine learning in Python. the Journal of machine Learning research 12, 2825–2830. [Google Scholar]
- 131.Bates D., Mächler M., Bolker B., and Walker S. (2015). Fitting Linear Mixed-Effects Models Using lme4. Journal of Statistical Software 67, 1–48. 10.18637/jss.v067.i01. [DOI] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.