Loading metrics

Open Access

Peer-reviewed

Research Article

Relating Structure and Function in the Human Brain: Relative Contributions of Anatomy, Stationary Dynamics, and Non-stationarities

* E-mail: [email protected]

Affiliation Laboratoire d'Imagerie Fonctionnelle, UMR678, Inserm/UPMC Univ Paris 06, Paris, France

  • Arnaud Messé, 
  • David Rudrauf, 
  • Habib Benali, 
  • Guillaume Marrelec

PLOS

  • Published: March 20, 2014
  • https://doi.org/10.1371/journal.pcbi.1003530
  • Reader Comments

Figure 1

Investigating the relationship between brain structure and function is a central endeavor for neuroscience research. Yet, the mechanisms shaping this relationship largely remain to be elucidated and are highly debated. In particular, the existence and relative contributions of anatomical constraints and dynamical physiological mechanisms of different types remain to be established. We addressed this issue by systematically comparing functional connectivity (FC) from resting-state functional magnetic resonance imaging data with simulations from increasingly complex computational models, and by manipulating anatomical connectivity obtained from fiber tractography based on diffusion-weighted imaging. We hypothesized that FC reflects the interplay of at least three types of components: (i) a backbone of anatomical connectivity, (ii) a stationary dynamical regime directly driven by the underlying anatomy, and (iii) other stationary and non-stationary dynamics not directly related to the anatomy. We showed that anatomical connectivity alone accounts for up to 15% of FC variance; that there is a stationary regime accounting for up to an additional 20% of variance and that this regime can be associated to a stationary FC; that a simple stationary model of FC better explains FC than more complex models; and that there is a large remaining variance (around 65%), which must contain the non-stationarities of FC evidenced in the literature. We also show that homotopic connections across cerebral hemispheres, which are typically improperly estimated, play a strong role in shaping all aspects of FC, notably indirect connections and the topographic organization of brain networks.

Author Summary

By analogy with the road network, the human brain is defined both by its anatomy (the ‘roads’), that is, the way neurons are shaped, clustered together and connected to each others and its dynamics (the ‘traffic’): electrical and chemical signals of various types, shapes and strength constantly propagate through the brain to support its sensorimotor and cognitive functions, its capacity to learn and adapt to disease, and to create consciousness. While anatomy and dynamics are organically intertwined (anatomy contributes to shape dynamics), the nature and strength of this relation remain largely mysterious. Various hypotheses have been proposed and tested using modern neuroimaging techniques combined with mathematical models of brain activity. In this study, we demonstrate the existence (and quantify the contribution) of a dynamical regime in the brain, coined ‘stationary’, that appears to be largely induced and shaped by the underlying anatomy. We also reveal the critical importance of specific anatomical connections in shaping the global anatomo-functional structure of this dynamical regime, notably connections between hemispheres.

Citation: Messé A, Rudrauf D, Benali H, Marrelec G (2014) Relating Structure and Function in the Human Brain: Relative Contributions of Anatomy, Stationary Dynamics, and Non-stationarities. PLoS Comput Biol 10(3): e1003530. https://doi.org/10.1371/journal.pcbi.1003530

Editor: Claus C. Hilgetag, Hamburg University, Germany

Received: August 14, 2013; Accepted: February 8, 2014; Published: March 20, 2014

Copyright: © 2014 Messé et al. This is an open-access article distributed under the terms of the Creative Commons Attribution License , which permits unrestricted use, distribution, and reproduction in any medium, provided the original author and source are credited.

Funding: This work is supported by the Inserm and the University Pierre et Marie Curie (Paris, France). The funders had no role in study design, data collection and analysis, decision to publish, or preparation of the manuscript.

Competing interests: The authors have declared that no competing interests exist.

Introduction

Coherent behavior and cognition involve synergies between neuronal populations in interaction [1] – [3] . Even at rest, in the absence of direct environmental stimulations, these interactions drive the synchronization of spontaneous activity across brain systems, shedding light on the large-scale anatomo-functional organization of the brain [4] . The study of such patterns of synchronization has known important developments due to recent methodological advances in brain imaging data acquisition and analysis. These advances have enabled investigators to estimate interactions in the brain by measuring functional connectivity (FC) from resting-state functional MRI (rs-fMRI). Analyses of FC at rest have supported the hypothesis that the brain is spatially organized into large-scale intrinsic networks [5] – [7] , e.g. the so-called resting-state networks [8] , [9] , such as the default mode network, which have been linked to central integrative cognitive functions [10] – [13] . The study of large-scale intrinsic networks from rs-fMRI has become a central and active area for neuroscience research. However, the mechanisms and factors driving FC, as well as their relative contribution to empirical data, are still highly debated [14] and remain to be elucidated.

Theoretical rationale and empirical findings support the hypothesis that FC is driven and shaped by structural connectivity (SC) between brain systems, i.e., by the actual bundles of white matter fiber connecting neurons [15] . As a first approximation, SC can be inferred from fiber tractography based on diffusion-weighted imaging (DWI) [16] – [19] . A recent study [20] , which focused on a small subset of robustly estimated structural connections, demonstrated the existence of a statistical, yet complex, correspondence between FC and specific features of SC (e.g., low vs. high fiber density, short vs. long fibers, intra vs. interhemispheric connections). However, a large part of FC cannot be explained by SC alone [21] . There appears that FC is the result of at least two main contributing factors: (i) the underlying anatomical structure of connectivity, and (ii) the dynamics of neuronal populations emerging from their physiology [3] . A key issue is to better understand the relative contributions of these two components to FC. Besides, recent studies using windowed analyses have suggested that FC estimated over an entire acquisition session (referred to as ‘stationary FC’ in the literature) breaks down into a variety of reliable correlation patterns (also referred to as ‘dynamic FC’ or ‘non-stationarities’) when estimated over short time windows (30 s) [14] , [22] . The authors advocated that FC estimated over short time windows (or windowed FC, for short) mostly reflects recurrent transitory patterns that are aggregated when estimating FC over a whole session. They further suggested that whole-session FC may only be an epiphenomenon without clear physiological underpinning, and not the reflection of an actual process with stationary FC [14] . This perspective remains to be reconciled with the fact that whole-session FC has been found to be highly reproducible, functionally meaningful and a useful biomarker in many pathological contexts [23] , [24] . Note that, in the recent literature of fMRI data analysis, stationarity implicitely refers to a stationary FC (i.e., the invariance of FC over time), to be contrasted with the more general notion of (strong) stationarity, where a model or process is stationary if its parameters remain constant over time [25] , [26] . SC being temporally stable at the scale of a whole resting state fMRI session (typically 10 min), we could expect SC to drive a stationary process (in the strong sense). Since SC is furthermore expected to drive FC, we can hypothesize that this stationary process contributes to generate a stationary FC.

In order to bring together the structural and dynamical components underlying FC, some studies have used computational models that incorporate SC together with biophysical models of neuronal activity to generate coherent brain dynamics [27] – [32] . This approach has yielded promising results for the understanding of the relationship between structure and function [17] , [33] , [34] . Here, we used a testbed of well-established generative models simulating neuronal dynamics combined with empirical measures, to investigate the relative contributions of anatomical connections, stationary dynamics, and non-stationarities to the emergence of empirical functional connectivity. In particular, we considered the following hypotheses: (H1) part of FC directly reflects SC; (H2) models of physiological mechanisms added to SC increase predictive power all the more as they are complex; (H3) part of the variance of FC that is unexplained by models is due to issues in the estimation of SC, e.g., problems with measuring homotopic connections; (H4) there is an actual stationary process reflected in whole-session FC that is not merely an artifact but substantially reflects the driving of the dynamics by SC.

In order to test these hypotheses and estimate the relative contribution of anatomy, stationary dynamics and non-stationarities to FC, we relied on the following approach. After T 1 -weighted MRI based parcellation of the brain ( N  = 160 regions), SC was estimated using the proportion of white matter fibers connecting pairs of regions, based on probabilistic tractography of DWI data [35] . FC was measured on rs-fMRI data using Pearson correlation between the time courses of brain regions. We quantified the correlation between SC alone and FC as a reference, and also fed SC to generative neurodynamical models of increasing complexity: a spatial autoregressive (SAR) model [36] , analytic models with or without conduction delays [28] – [31] , [37] , and biologically constrained models [29] , [32] . Importantly, all these models were used in their stationary regime in the strong sense, since their parameters were not changed during the simulations. Of these models, only the SAR is explicitely associated with a stationary FC; other, more complex models, generate dynamics that are compatible with a non-stationary FC. We computed FC from data simulated by these models and compared the results to empirical FC. For each model, performance was quantified using predictive power [29] , for each subject as well as on the ‘average subject’ (obtained by averaging SC and empirical FC across subjects). Values for the model parameters were based on the literature, except for the structural coupling strength that was optimized in order to maximize each model's performance.

Predictive power of models

In agreement with H1, SC explained a significant amount of the variance of whole-session FC for all subjects, as did all generative models (permutation test, p <0.05 corrected) ( Figure 1 , panel A). Generative models predicted FC better than SC alone (paired permutation test, p <0.05 corrected). Predictive power obtained with the average subject ranged from 0.32 for SC alone to 0.43 for the SAR model ( Table 1 ). For a given model, predictive power was reproducible across subjects. Contrary to our hypothesis H2, generative models had similar performance, and complexity was not predictive of performance. The results remained unchanged when no global signal regression was applied ( Figure S1 ). Also, findings were found to be similar for SC alone and the SAR model at finer spatial scales ( N  = 461 and N  = 825 regions, Figure S2 ) and consistent with a replication dataset ( Figure S3 ). Most importantly, a large part of the variance ( R 2 ) in the empirical data (at least 82%) remained unexplained by this first round of simulations.

thumbnail

  • PPT PowerPoint slide
  • PNG larger image
  • TIFF original image

(A) Predictive power for all connections and when restricted to intra/interhemispheric, direct/indirect connections. For each type of connections and each model, we represented the individual predictive powers (bar chart representing means and associated standard deviations), as well as the predictive power for the average subject computed using the original SC (diamonds), or after adding homotopic connections (circles). Of note, SC alone has no predictive power (zero) for the subset of indirect connections, by definition. (B) Patterns of SC, empirical FC and FC simulated from the SAR model for the average subject and associated scatter plot of simulated versus empirical FC (solid line represents perfect match). SARh stands for the SAR model with added homotopic connections. Matrices were rearranged such that network structure is highlighted. Homologous regions were arranged symmetrically with respect to the center of the matrix; for instance, the first and last regions are homologous. (C) Similarity of functional brain networks across subjects (bar chart with means and associated standard deviations), for the average subject (diamonds), and when adding homotopic connections (circles) (left). Network maps for the average subject and empirical FC, as well as for FC simulated using either the SAR model with original SC or the SARh.

https://doi.org/10.1371/journal.pcbi.1003530.g001

thumbnail

https://doi.org/10.1371/journal.pcbi.1003530.t001

Role of homotopic connections

We reasoned (see hypothesis H3) that part of the unexplained variance could reflect issues with the estimation of SC from DWI, which can be expected because of limitations in current fiber tracking algorithms and the problem of crossing fibers [38] . We know for instance that many fibers passing through the corpus callosum are poorly estimated in diffusion imaging, in particular those connecting more lateral parts of the cerebral cortex [39] . Yet, the corpus callosum is the main interhemispheric commissure of the mammal brain, see [40] . It systematically connects homologous sectors of the cerebral cortex across the two hemispheres in a topographically organized manner, with an antero-posterior gradient, through a system of myelinated homotopic fibers or ‘homotopic connections’. The hypothesis of an impact of SC estimation problems on FC unexplained variance was supported by the observation that, in our results, intrahemispheric connections yielded on average a much higher predictive power (e.g., 0.59 for the SAR model) than interhemispheric connections (0.16 for the SAR model).

In order to further test the role of white matter connections in driving FC, we artificially set all homotopic connections to a constant SC value (0.5) for the average subject and reran all simulations. As a result, the predictive power strongly increased for all models ( Figure 1 , panels A and B), ranging from 0.39 for SC alone to 0.61 for the SAR model ( Table 1 ). Thus the variance unexplained (1- R 2 ) was reduced to 63%. Moreover, predictive power for intra and interhemispheric connections became equivalent (0.60 and 0.62, respectively). Interestingly, adding homotopic connections also led to a substantial increase in predictive power for indirect connections, that is, pairs of regions for which SC is zero (increasing from 0.07 to 0.45). The effect of adding interhemispheric anatomical connections on increasing predictive power was highly specific to homotopic connections. When applying the SAR model to the SC matrix with added homotopic connections and randomly permuting (10 000 permutations) the 80 corresponding interhemispheric connections (one region in one hemisphere was connected to one and only one region in the other hemisphere), the predictive power strongly decreased, even compared to results with the original SC ( Figure 2 , panel A). Moreover, we further assessed the specificity of this result by systematically manipulating SC. In three different simulations, we randomly removed, added, and permuted structural connections (10 000 times). In all cases, the predictive power decreased as a function of the proportion of connections manipulated ( Figure 2 , panel B). Moreover, changes induced by these manipulations remained small (<0.05), far below the changes that we were able to induce by adding homotopic connections. All in all, these results suggest that homotopic connections play a key role in shaping the network dynamics, in a complex and non-trivial manner.

thumbnail

(A) Predictive power of the SAR model with original SC (green), when adding homotopic connections (‘SARh’, red), or with shuffled homotopic connections (black). (B) Predictive power of the SAR model with original SC (red) and when SC values were randomly permuted, removed or added (from left to right). For each graph, predictive power was quantified as a function of the percentage of connections manipulated.

https://doi.org/10.1371/journal.pcbi.1003530.g002

Predicting functional brain networks

Beyond predicting the overall pattern of FC, we also assessed whether models could predict the empirical organization of FC into a set of intrinsic networks. Connectivity matrices were clustered into groups of non-overlapping brain regions showing high within-group correlation and low between-group correlation, and the resulting partitions into functional brain networks were compared between empirical and simulated FC using the adjusted Rand index (see Methods ). Again, the SAR model tended to perform best among all computational models ( Figure 1 , panel C).

Without adding homotopic connections in the SC matrix, the simulated networks highly differed from the empirical networks. In particular, most networks were found to be lateralized. After adding homotopic connections, the resemblence between simulated and empirical networks greatly improved. Networks were more often bilateral and overall consistent with the topography of empirical functional networks, including somatosensory, motor, visual, and associative networks. High FC between the amygdala and ventral-lateral sectors of the prefrontal cortex was also correctly predicted by the simulations. There were nevertheless some notable differences. First, the clustering of empirical FC yielded a long-range fronto-parieto-temporal association network ( Figure 1 , panel C, cyan) that was not observed in simulated FC as such. Second, a parieto-temporal cluster ( Figure 1 , panel C, red), which was associated with thalamo-striatal networks, was predicted by simulations but was not present in the empirical data. Third, a cluster encompassing the entire cingulate gyrus and precuneus ( Figure 1 , panel C, green) was predicted by simulations but was broken down into more clusters in the empirical data.

Stationary FC, non-stationary FC, and non-stationarities

The results above show that SC plays a causal role in FC, but one can still wonder what aspects of the underlying dynamics are the most directly related to this influence. A hypothesis is that SC, in combination with stable physiological processes (e.g., overall gain in synaptic transmission), drives a stationary regime of the dynamics. This hypothesis is supported by the finding that all models tested in this study, which were used in a stationary regime (in the strong sense), could explain significantly more variance than SC alone. Furthermore, the fact that the SAR could predict FC significantly better than all other models is evidence that this stationary regime is associated with stationary FC (paired permutation test, p <0.05 corrected).

But, clearly, many variations in the dynamical patterns of brain activity, be it in the process of spontaneous cognition, physiological regulation, or context-dependent changes, cannot be expected to be associated with a purely stationary FC. Modeling how the brain dynamics deal with endogenous and environmental contexts should require more complex models, either stationary or non-stationary, that are able to generate non-stationary (i.e., time-varying) patterns of FC. Given that at best 37% of the variance could be explained by the model of a purely stationary FC (the SAR), we can wonder why the models of higher complexity used in our simulation testbed did not perform better in predicting FC. One possible hypothesis is that the SAR model was favored in the simulations, because we estimated FC over about 10 minutes of actual brain dynamics. In such configuration, we can imagine that the non-stationarities of FC cancel out, the estimation effectively keeping the stationary part of FC. We thus wondered whether the more complex models would better perform when non-stationary FC had the potential of being more strongly reflected in the data. We approached this question by computing predictive power on windowed FC as a function of the length of the time-window used [22] , for all possible time-windows over which FC could be estimated and for all models. We also investigated the effect of simulation duration (see Methods ). We found that the relative performance of more complex models was still lower than that of the SAR model ( Figures 3 and S4 ). The average predictive power was lower for shorter time-windows and increased towards a limit for longer time-windows. The SAR model behaved like an ‘upper-bound’ for predictive power. The performance of all other models, irrespective of the size of the time-window, was between that of SC alone and that of the SAR model.

thumbnail

Predictive power as a function of time-window length across subjects (left) and of duration of simulated runs on the average subject (right). For color code see Figure 1 .

https://doi.org/10.1371/journal.pcbi.1003530.g003

A straightforward explanation is that the non-stationary patterns of FC, as generated by the simulation models, did not match the non-stationary patterns of the empirical FC as they unfolded during the acquisition in the brain of the participants. Context-dependent and transient dynamics are likely to be missed by models of the dynamics that cannot be contextually constrained in the absence of further information. It is thus difficult to infer how much of the 63% of unexplained variance remaining in whole-session FC actually reflect physiologically meaningful non-stationary FC, and more broadly, non-stationary dynamics.

In the present study, we investigated the respective contributions of anatomical connections, stationary dynamics, and non-stationarities to the emergence of empirical functional connectivity. We compared the performance of computational models in modeling FC and manipulated SC in order to analyze the impact of SC on FC, with and without the filter of combined physiological models of the dynamics.

The importance of white matter fiber pathways in shaping functional brain networks is a known fact, for a review, see [15] , [17] , [21] , [23] . Previous modeling studies have supported the importance of the underlying anatomical connections, i.e., SC, in shaping functional relationships among brain systems [16] , [41] , [42] . In agreement with our hypothesis H1, we showed that functional connectivity could at least in part be explained by structural connectivity alone. Adding homotopic connections in the matrix of SC, we found a slight increase in explained variance when considering the prediction of whole-session FC from SC alone (+4% of explained variance). In agreement with H2, adding models of physiological interactions above and beyond SC alone increased the explained variance in whole-session FC, by 8% for the best performing model, the SAR model, when no homotopic connections were added, and by 22% when homotopic connections were added. This latter fact, which strongly supports H3, suggests a complex interplay between anatomy as reflected by SC and physiological mechanisms in generating FC. This impact of SC manipulations on predicted FC pertained not only to direct but also to indirect connections. For indirect connections, whole-session FC was much better predicted after adding homotopic connections to SC than before adding them (0.45 versus 0.07 in predictive power). The problem of limited predictive power for FC based on SC when considering indirect connections has puzzled the field [43] . For this reasons many studies only assess the performance of models on direct connections. Here, we showed that a major factor in driving FC for indirect anatomical connections (+20% in explained variance) is the interplay between a subset of anatomical connections, i.e., homotopic connections (which are typically underestimated by DWI), and physiological parameters that generate the dynamics underlying FC, themselves conditioned by the possible interactions defined by SC.

Contrary to our expectation (see hypothesis H2), all models tended to perform similarly, irrespective of model complexity. The best performing model in most cases was the SAR model, a model of stationary FC driven by SC, with 63% of the variance remaining unexplained. It is likely that, above and beyond problems with the estimation of SC from DWI, and other incompressible sources of irrelevant noise, much of the unexplained variance in FC relates to non-stationary patterns in FC, and more generally to non-stationarities in the strong-sense. Such non-stationarities are difficult to model in the experimentally unconstrained resting-state and in the absence of further information regarding the specific parameters shaping FC. Irrespective of their complexity, computational models are only capable of generating prototypical brain activities, and not the subject-dependent activity that took place in the brain of the participants during scanning. The scientific necessity of modeling brain dynamics is hindered by such uncertainty and it will be a challenge to find solutions to approach this problem [26] , [44] . Even though one objective for neuroscience is to propose generative models that are capable of generating detailed neuronal dynamics, generative models cannot be informed by this unknown context and, as a consequence, cannot generate context-dependent activity in a manner that would be predictive of empirical data, in the absence of additional measures and experimental controls. Nevertheless, and perhaps for that very reason, the study of non-stationarity in FC should become of central interest for the field, as such non-stationarities could explain much of FC (up to 63% according to our simulation results), and thus reflect critical mechanisms for neurocognitive processing.

In the absence of adequate modeling principles, determining the precise contribution of non-stationarities to the unexplained variance in FC is impossible, as other confounding sources of unexplained variance are expected. As we showed, even naive manipulations aimed at estimating the impact of the known errors in DWI-based reconstruction of homotopic connections showed that such errors could cause 20% of the unexplained variance in predicting empirical FC. How DWI and fiber tracking should be used for an optimal estimation of structural connectivity is still a topic of intense debates [45] – [48] . It is likely that part of the unexplained variance in predicting FC will be reduced as better estimates of SC become available.

The model showing the best results, the SAR model, explicitly modeled a stationary process with a stationary FC. In line with our hypothesis H4, empirical FC is likely to incorporate stationary components driven by SC. Further knowledge about this stationary process might be gained by analyzing FC computed over much longer periods of time than is commonly performed (e.g., hours versus minutes). This stationary process is itself likely to be only locally stationary, as it might be expected that slow physiological cycles, from nycthemeral cycles to hormonal cycles, development, learning and aging, will modify the parameters controlling it.

In the present study, we did not take into account the statistical fluctuations induced by the fact that the time series were of finite length. Such a finiteness entails that even a model that is stationary in the strong sense could generate sample moments that fluctuate over time. For instance, the sample sum of square of a multivariate normal model with covariance matrix Σ computed from time series of size N is not equal to N Σ but is Wishart distributed with N -1 degrees of freedom and scale matrix Σ. This phenomenon will artificially increase the part of variance that cannot be accounted for by stationary models and, hence, play against stationary models. Since it is conversely very unlikely for a non-stationary model to generate sample moments that are constant over time, statistical fluctuations cannot at the same time artificially increase the part of variance that can be accounted for by stationary models. As a consequence, not considering these statistical fluctuations made us underestimate the part of variance that can be accounted for by models that are stationary in the strong sense. In other words, our estimate of the part of variance accounted for by a stationary model is a lower bound for the true value. We can therefore be confident that taking statistical fluctuations into account will only strengthen H4.

Our goal here was to investigate how current generative models of brain acticity fare in predicting the relationship between structure and function. The complexity of some of these models was such that the simulations included here were only possible thanks to a computer cluster. The behavior of all these models depends on the values of some parameters and, in the present study, we set these parameters in agreement with the literature. In what measure this choice affects how well models predict FC is unclear. Yet a full investigation of this issue remains beyond the scope of this study, since parameter optimization through extensive exploration of the parameter space for all models is at this stage unrealistic. Nevertheless, in order to get a sense of the sensitivity of our results to parameter values in a way that is compatible with the computational power available, we explored the behavior of the Fitzhugh-Nagumo, Wilson and Kuramoto models over a subset of the parameter space (see Figures S5 and S6 ). We found that parameter values had little influence on predictive power, which, in all cases, remained below that of the SAR, the simplest model tested.

We formulated H2 to test for the existence of a relationship between complexity and realism in the models that we used. Indeed, there should exist a very tight connection between the two, since the more complex generative models in our study have been designed to take biophysical mechanisms into account, with parameters that are physiologically relevant and values often chosen based on prior experimental results. Now, realism usually comes at the cost of complexity. As a consequence, it is often (implicitely) assumed that, among the models we selected, the more complex a model is, the more realistic it will also be and the better it will fit the data. This is the reason why we stated H2, based on such rationale inspired from the literature, in order to put such hypothesis to the test. The results show that for the models we used, with their sets of parameters, an increase in complexity was not associated with an increase in performance. This suggests that, for these models, complexity and realism are not quite as tightly connected as expected.

Given that the SAR model is the only model that does not include a step of hemodynamic modeling (Balloon-Windkessel), it cannot be ruled out that the superiority of the SAR reflects issues with this step. In order to check that this is not the case, we computed predictive power for all models without the hemodynamic model. The predictive power was largely insensitive to the presence of the hemodynamic model (see Figure S7 ). In particular, the SAR model remained overall an upper bound in terms of predictive power.

Finally, we should note that we relied on a definition of SC restricted to the white matter compartment. Although this is standard in the field, in reality, local intrinsic SC exists in the gray matter. However, current models generally make prior assumptions about such SC. Moreover, intrinsic SC currently remains impossible to measure reliably for the entire brain.

In spite of the complexity of the problems and the limitations of current modeling approaches, computational modeling of large-scale brain dynamics remains an essential scientific endeavor. It is key to better understand generative mechanisms and make progress in brain physiology, physiopathology and, more generally, theoretical neuroscience. It is also central to the endeavor of searching for accurate and meaningful biomarkers in aging and disease [49] . Moreover, computational modeling of FC opens the possibility of making inference on specific biophysical parameters, including inference about the underlying anatomical connectivity itself. In spite of their limited predictive powers, simpler models can be useful in this context. The SAR model, introduced in [36] , may appear well-suited to model essential stationary aspects of the generative mechanisms of FC. One interest of such a simple and analytically tractable model is that, beyond its very low computational burden, it could be the basis for straightforward estimation of the model parameters that can be used to compare clinical populations, and could constitute a potentially important biomarker of disease.

Ethics statement

All participants gave written informed consent and the protocol was approved by the local Ethics Committee of the Pitié-Salpêtrière Hospital (number: 44-08; Paris, France).

Twenty-one right-handed healthy volunteers were recruited within local community (11 males, mean age 22±2.4 years). Data were acquired using a 3 T Siemens Trio TIM MRI scanner (CENIR, Paris, France). For acquisition and preprocessing details, see Text S1 . For each subject, the preprocessing yielded three matrices: one of SC, one with the average fiber lengths, and one of empirical FC. These matrices were also averaged across subjects (‘average subject’).

Simulations

We used eight generative models with various levels of complexity: the SAR model, a purely spatial model with no dynamics that expresses BOLD fluctuations within one region as a linear combination of the fluctuations in other regions; the Wilson-Cowan system, a model expressing excitatory and inhibitory neuronal populations activity; the two rate models (with or without conduction delays), simplified versions of the Wilson-Cowan system obtained by considering exclusively the excitatory population; the Kuramoto model, which simulates neuronal activity using oscillators; the Fitzhugh-Nagumo model, which aims at reproducing complex behaviors such as those observed in conductance-based models; the neural-mass model, also based on conductance and with strong biophysiological constraints; and finally, the model of spiking neurons, the most constrained model in the current study which models neuron populations as attractors. For more details, see Text S2 .

All models took an SC matrix as input, and all but the SAR were taken as models of neuronal (rather than BOLD) activity. Simulated fMRI BOLD signal was obtained from simulated neuronal activity by means of the Balloon-Windkessel hemodynamic model [50] , [51] . Global mean signal was then regressed out from each region's time series. Finally, simulated FC was computed as Pearson correlation between simulated time series. For the SAR model, we directly computed simulated FC from the analytical expression of the covariance, see Equation (2) in Text S2 . All models had a parameter that represents the coupling strength between regions. This parameter was optimized separately for each model on the average subject to limit computational burden ( Text S2 ). After optimization, we generated three runs of 8 min BOLD activity and averaged the corresponding FCs to obtain the simulated FC for each dynamical model and each subject. For the average subject, simulated FC was obtained by feeding the average SC matrix to the different models.

Performance

Modeling performance was assessed using predictive power and similarity of spatial patterns. Predictive power was quantified for each subject and for the average subject by means of Pearson correlation between simulated and empirical FC [29] . Regarding the similarity of functional brain networks, SC, empirical FC and simulated FC were decomposed into 10 networks using agglomerative hierarchical clustering and generalized Ward criterion [52] . The resulting networks from SC and simulated FC were compared to the ones resulting from empirical FC using the adjusted Rand index [53] , [54] . The Rand index quantifies the similarity between two partitions of the brain into networks by computing the proportion of pairs of regions for which the two partitions are consistent (i.e., they are either in the same network for both partitions, or in a different network for both partitions). The adjustment accounts for the level of similarity that would be expected by chance only.

Analysis of dynamics

Empirical and simulated windowed FC were computed on individual subjects using sliding time-windows (increment of 20 s) of varying length (from 20 to 420 s by step of 20 s). Predictive power was computed as the correlation between any pair of time-windows of equal length corresponding to simulated and empirical windowed FC, respectively. This approach was only applied to the dynamical models; for SC alone and the SAR model, simulated FC remained, by definition, constant through time and, as a consequence, windowed FC was equaled to whole-session FC. The influence of simulated run duration on predictive power was also investigated. For each model, three runs of one hour were simulated on the average subject. Predictive power was then computed as a function of simulated run duration. For the same reason as above, SC alone and the SAR model did not depend on simulation duration.

Supporting Information

Performance of computational models when no global signal regression was performed. (A) Predictive power for all connections and when restricted to intra/interhemispheric, direct/indirect connections. For each type of connections and each model, we represented the individual predictive powers (bar chart representing means and associated standard deviations), as well as the predictive power for the average subject computed using the original SC (diamonds), or after adding homotopic connections (circles). Of note, SC alone has no predictive power (zero) for the subset of indirect connections, by definition. (B) Patterns of SC, empirical FC and FC simulated from the SAR model for the average subject and associated scatter plot of simulated versus empirical FC (solid line represents perfect match). SARh stands for the SAR model with added homotopic connections. Matrices were rearranged such that network structure is highlighted. Homologous regions were arranged symmetrically with respect to the center of the matrix; for instance, the first and last regions are homologous. (C) Similarity of functional brain networks across subjects (bar chart with means and associated standard deviations), for the average subject (diamonds), and when adding homotopic connections (circles) (left). Network maps for the average subject and empirical FC, as well as for FC simulated using either the SAR model with original SC or the SARh.

https://doi.org/10.1371/journal.pcbi.1003530.s001

Performance of SC alone and the SAR model at finer spatial scales. Predictive power for all connections and when restricted to intra/interhemispheric, direct/indirect connections. For each type of connections and each model, we represented the individual predictive powers (bar chart representing mean and associated standard deviation), as well as the predictive power of the average subject computed using the original SC (diamonds), or after adding homotopic connections (circles).

https://doi.org/10.1371/journal.pcbi.1003530.s002

Performance of computational models on the replication dataset. The replication dataset was from the study of Hagmann and colleagues [55] . Brain network was defined at low anatomical granularity (N = 66 regions), and connectivity measures were averaged over five healthy volunteer subjects. (A) Predictive power for all connections and when restricted to intra/interhemispheric, direct/indirect connections. For each type of connections and each model, we represented the individual predictive powers (bar chart representing means and associated standard deviations), as well as the predictive power for the average subject computed using the original SC (diamonds), or after adding homotopic connections (circles). Of note, SC alone has no predictive power (zero) for the subset of indirect connections, by definition. (B) Patterns of SC, empirical FC and FC simulated from the SAR model for the average subject and associated scatter plot of simulated versus empirical FC (solid line represents perfect match). SARh stands for the SAR model with added homotopic connections. Matrices were rearranged such that network structure is highlighted. Homologous regions were arranged symmetrically with respect to the center of the matrix; for instance, the first and last regions are homologous. (C) Similarity of functional brain networks across subjects (bar chart with means and associated standard deviations), for the average subject (diamonds), and when adding homotopic connections (circles) (left). Network maps for the average subject and empirical FC, as well as for FC simulated using either the SAR model with original SC or the SARh.

https://doi.org/10.1371/journal.pcbi.1003530.s003

Effect of time on performance. Predictive power of computational models as a function of the time-window length for each subject (graphs) and model (color).

https://doi.org/10.1371/journal.pcbi.1003530.s004

Exploration of the parameter space for the Fitzhugh-Nagumo model. (Left) Phase diagrams (i.e., x - y plane) for an uncoupled model ( k  = 0) over various parameter values of α and β . The model operate mostly in an oscillatory regime for the range of parameter values investigated. (Right) Predictive power as a function of α and β . The black dot represents the parameter set used in our simulations, while the black square corresponds to the values from [28] . The values used in our simulations gave rise to higher predictive power than the parameters values from [28] . In any case, for the range of parameters considered, the predictive power always remained lower than that obtained with a SAR model.

https://doi.org/10.1371/journal.pcbi.1003530.s005

Effect of velocity on predictive power. Predictive power as a function of the coupling strength and velocity values in generative models. Black dots represent values used for subsequent simulations. These simulations show that the predictive power is little influenced by velocity. In any case, for the range of parameters considered, the predictive power also always remained lower than that obtained with a SAR model.

https://doi.org/10.1371/journal.pcbi.1003530.s006

Effect of the hemodynamic model. Predictive power for all connections and when restricted to intra/interhemispheric, direct/indirect connections. For each type of connections and each model, we represented the predictive power for the average subject computed using the BOLD signal (diamonds, solid line) or using the neuronal activity (circles, dashed line). Of note, the prediction differs slightly from that of the Figure 1 due to the stochastic component of most models at each run.

https://doi.org/10.1371/journal.pcbi.1003530.s007

Data and preprocessing.

https://doi.org/10.1371/journal.pcbi.1003530.s008

Computational models.

https://doi.org/10.1371/journal.pcbi.1003530.s009

Acknowledgments

The authors are thankful to Olaf Sporns (Department of Psychology, Princeton University, Princeton, USA) and Christopher J. Honey (Department of Psychological and Brain Sciences, Indiana University, Bloomington, USA) for providing the neural-mass model; to Gustavo Deco, Étienne Hugues and Joanna Cabral (Computational Neuroscience Group, Department of Technology, Universitat pompeu Fabra, Barcelona, Spain) for providing the Kuramoto and rate models as well as the spike model; and to Olaf Sporns and Patric Hagmann (Department of Radiology, University Hospital Center and University of Lausanne, Lausanne, Switzerland) for sharing their data for replication. We would also like to thank them for fruitful discussions. The authors are grateful to Stéphane Lehéricy and his team (Center for Neuroimaging Research, Paris, France) for providing them with the data, and especially to Romain Valabrègue for his help in handling coarse-grained distributed parallelization of computational tasks.

Author Contributions

Conceived and designed the experiments: AM DR HB GM. Performed the experiments: AM DR GM. Analyzed the data: AM. Wrote the paper: AM DR HB GM.

  • View Article
  • Google Scholar
  • 40. Schmahmann JD, Pandya DN (2006) Fiber pathways of the brain. Oxford University Press.
  • 52. Batageli V (1988) Generalized ward and related clustering problems. In: Bock HH, editor. Classification and Related Methods of Data Analysis. North-Holland. pp. 67–74.

Information

  • Author Services

Initiatives

You are accessing a machine-readable page. In order to be human-readable, please install an RSS reader.

All articles published by MDPI are made immediately available worldwide under an open access license. No special permission is required to reuse all or part of the article published by MDPI, including figures and tables. For articles published under an open access Creative Common CC BY license, any part of the article may be reused without permission provided that the original article is clearly cited. For more information, please refer to https://www.mdpi.com/openaccess .

Feature papers represent the most advanced research with significant potential for high impact in the field. A Feature Paper should be a substantial original Article that involves several techniques or approaches, provides an outlook for future research directions and describes possible research applications.

Feature papers are submitted upon individual invitation or recommendation by the scientific editors and must receive positive feedback from the reviewers.

Editor’s Choice articles are based on recommendations by the scientific editors of MDPI journals from around the world. Editors select a small number of articles recently published in the journal that they believe will be particularly interesting to readers, or important in the respective research area. The aim is to provide a snapshot of some of the most exciting work published in the various research areas of the journal.

Original Submission Date Received: .

  • Active Journals
  • Find a Journal
  • Proceedings Series
  • For Authors
  • For Reviewers
  • For Editors
  • For Librarians
  • For Publishers
  • For Societies
  • For Conference Organizers
  • Open Access Policy
  • Institutional Open Access Program
  • Special Issues Guidelines
  • Editorial Process
  • Research and Publication Ethics
  • Article Processing Charges
  • Testimonials
  • Preprints.org
  • SciProfiles
  • Encyclopedia

life-logo

Article Menu

research paper on brain anatomy

  • Subscribe SciFeed
  • PubMed/Medline
  • Google Scholar
  • on Google Scholar
  • Table of Contents

Find support for a specific problem in the support section of our website.

Please let us know what you think of our products and services.

Visit our dedicated information section to learn more about MDPI.

JSmol Viewer

Brain structure and function: insights from chemical neuroanatomy.

research paper on brain anatomy

1. General Premises

2. from the contributions of golgi and cajal to modern chemical neuroanatomy: a brief survey.

  • Private channel: physically delimited pathway between two nodes of the network
  • Diffuse channel: the whole available space between the network nodes is potentially used to exchange signals
  • Reserved signal: Signal needing a specific “decoder” in order to be decrypted. Neurotransmitters and, more generally, signals using specific receptor systems are of this type
  • Broadcast signal: “Public” signal, i.e., interpreted by all the elements that it can reach. Physical processes (e.g., pressure waves) or membrane permeable molecules (e.g., oxygen) are of this type
  • Electrical signals from the pre-synaptic side can affect the post-synaptic side by means of induction;
  • Electrical signals can be conducted by the extracellular fluid (electrotonic currents);
  • A chemical mediator (neurotransmitter) can cross the synaptic cleft;
  • Transient connection can take place between the pre-synaptic and post-synaptic neuron and also via the extracellular matrix surrounding the synaptic contact; the matrix is part of the extracellular molecular network, and affects pre- and post-synaptic morpho-functional aspects of some synaptic contacts [ 94 ].

3. Experimental Contributions to Investigations of the Morpho-Functional Organization of Brain Networks at Different Levels of Miniaturization

3.1. the “mismatch” in several histochemical images between the nerve terminals, and hence the neurotransmitter stores, and their respective decoding receptors: a basic datum in the proposal of non-synaptic transmission, i.e., volume transmission, 3.1.1. types of vt signals, 3.1.2. pathways of vt-signal migration, 3.1.3. energy gradients for vt-signal migration.

  • Concentration Gradients (ref. [ 104 ] and References Therein)
  • Gradients of Electrical Potentials (for Charged Signals) (ref. [ 104 ] and References Therein)
  • Pressure Gradients (ref. [ 104 ] and References Therein)
  • Temperature Gradients (ref. [ 104 ]; See also [ 107 ])

3.1.4. Decoding Systems for VT-Signals

3.2. evidence of the existence of horizontal molecular networks at cell membrane levels and of the integrative role of gpcr aggregates.

  • The possible appearance of new binding sites or binding characteristics in each monomer (refs. [ 121 , 123 , 126 , 127 ]; see also [ 140 ]);
  • Different localization of the RM at plasma membrane levels, in comparison with the isolated monomers (e.g., preferential localization in the lipid rafts) [ 141 , 142 ];
  • Different turnover rate and desensitization of the monomers in the RM in comparison with the isolated GPCRs [ 143 , 144 , 145 ];
  • The possible existence of a “Hub Receptor” in an RM that is made up of three or more GPCRs. The Hub Receptor has been defined as the GPCR that can interact with multiple molecules, including receptors of the RM or membrane-associated proteins [ 18 , 144 , 145 , 146 , 147 , 148 , 149 ].

4. Future Investigations on Integrative Functions of the Brain

5. final comment: epistemological considerations.

  •    The Brain—is wider than the Sky—
  •    For—put them side by side—
  •    The one the other will contain
  •    With ease—and You—beside—.

Author Contributions

Institutional review board statement, informed consent statement, data availability statement, acknowledgments, conflicts of interest, list of abbreviations.

BHNBrain Hyper-Network
CNSCentral Nervous System
GPCRG Protein-Coupled Receptor
RMReceptor Mosaic
RRIReceptor–receptor interaction
VTVolume Transmission
WTWiring Transmission
  • Kuhn, T.S. The Structure of Scientific Revolutions , 1st ed.; University of Chicago Press: Chicago, IL, USA, 1962. [ Google Scholar ]
  • Kuhn, T.S. The Structure of Scientific Revolutions , 3rd ed.; University of Chicago Press: Chicago, IL, USA, 1996. [ Google Scholar ]
  • Björklund, A.; Hökfelt, T. Handbook of Chemical Neuroanatomy, Vol. 1: Methods in Chemical Neuroanatomy ; Elsevier Science Publishers: Amsterdam, The Netherlands, 1983; ISBN 0444902813. [ Google Scholar ]
  • Kuhar, M.J.; Unnerstall, J.R.; De Souza, E.B. Receptor mapping in neuropharmacology by autoradiography: Some technical problems. NIDA Res. Monogr. 1985 , 62 , 1–12. [ Google Scholar ] [ PubMed ]
  • Agnati, L.F.; Fuxe, K. Quantitative Neuroanatomy in Transmitter Research ; Macmillan Press Ltd.: London, UK, 1985. [ Google Scholar ]
  • Sporns, O. Discovering the Human Connectome ; MIT Press: Cambridge, MA, USA, 2012. [ Google Scholar ]
  • Krishnan, V.; Nestler, E.J. The molecular neurobiology of depression. Nature 2008 , 455 , 894–902. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Finnerup, N.B.; Kuner, R.; Jensen, T.S. Neuropathic pain: From mechanisms to treatment. Physiol. Rev. 2021 , 101 , 259–301. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Yang, M.; Yang, Z.; Wang, P.; Sun, Z. Current application and future directions of photobiomodulation in central nervous diseases. Neural Regen. Res. 2021 , 16 , 1177–1185. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • May, L.T.; Leach, K.; Sexton, P.M.; Christopoulos, A. Allosteric modulation of G protein-coupled receptors. Annu. Rev. Pharmacol. Toxicol. 2007 , 47 , 1–51. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Hauser, A.S.; Attwood, M.M.; Rask-Andersen, M.; Schiöth, H.B.; Gloriam, D.E. Trends in GPCR drug discovery: New agents, targets and indications. Nat. Rev. Drug Discov. 2017 , 16 , 829–842. [ Google Scholar ] [ CrossRef ]
  • Nussinov, R.; Tsai, C.J. Allostery in disease and in drug discovery. Cell 2013 , 153 , 293–305. [ Google Scholar ] [ CrossRef ]
  • Nickols, H.H.; Conn, P.J. Development of allosteric modulators of GPCRs for treatment of CNS disorders. Neurobiol. Dis. 2014 , 61 , 55–71. [ Google Scholar ] [ CrossRef ]
  • Agnati, L.F.; Baluška, F.; Barlow, P.W.; Guidolin, D. Mosaic, self-similarity logic and biological attraction principles: Three explanatory instruments in biology. Commun. Integr. Biol. 2009 , 2 , 552–563. [ Google Scholar ] [ CrossRef ]
  • Milligan, G. G protein-coupled receptor hetero-dimerization: Contribution to pharmacology and function. Br. J. Pharmacol. 2009 , 158 , 5–14. [ Google Scholar ] [ CrossRef ]
  • Agnati, L.F.; Guidolin, D.; Leo, G.; Carone, C.; Genedani, S.; Fuxe, K. Receptor-receptor interactions: A novel concept in brain integration. Prog. Neurobiol. 2010 , 90 , 157–175. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Lane, J.R.; Donthamsetti, P.; Shonberg, J.; Draper-Joyce, C.J.; Dentry, S.; Michino, M.; Shi, L.; López, L.; Scammells, P.J.; Capuano, B.; et al. A new mechanism of allostery in a G protein-coupled receptor dimer. Nat. Chem. Biol. 2014 , 10 , 745–752. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Farran, B. An update on the physiological and therapeutic relevance of GPCR oligomers. Pharmacol. Res. 2017 , 117 , 303–327. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Guidolin, D.; Marcoli, M.; Tortorella, C.; Maura, G.; Agnati, L.F. G protein-coupled receptor-receptor interactions give integrative dynamics to intercellular communication. Rev. Neurosci. 2018 , 29 , 703–726. [ Google Scholar ] [ CrossRef ]
  • Guidolin, D.; Marcoli, M.; Tortorella, C.; Maura, G.; Agnati, L.F. Receptor-Receptor Interactions as a Widespread Phenomenon: Novel Targets for Drug Development? Front. Endocrinol. 2019 , 10 , 53. [ Google Scholar ] [ CrossRef ]
  • Slosky, L.M.; Caron, M.G.; Barak, L.S. Biased Allosteric Modulators: New Frontiers in GPCR Drug Discovery. Trends Pharmacol. Sci. 2021 , 42 , 283–299. [ Google Scholar ] [ CrossRef ]
  • Agnati, L.F.; Genedani, S.; Leo, G.; Rivera, A.; Guidolin, D.; Fuxe, K. One century of progress in neuroscience founded on Golgi and Cajal’s outstanding experimental and theoretical contributions. Brain Res. Rev. 2007 , 55 , 167–189. [ Google Scholar ] [ CrossRef ]
  • Mazzarello, P. Golgi C: A Biography of the Founder of Modern Neuroscience ; Oxford University Press: New York, NY, USA, 2010. [ Google Scholar ]
  • Golgi, C. The Neuron Doctrine—Theory and Facts; Nobel Lecture. 11 December 1906. Available online: https://www.nobelprize.org/prizes/medicine/1906/golgi/lecture/ (accessed on 9 November 2022).
  • Cajal, S.R. The Structure and Connexions of Neurons; Nobel Lecture. 12 December 1906. Available online: https://www.nobelprize.org/prizes/medicine/1906/cajal/lecture/ (accessed on 9 November 2022).
  • Sherrington, C.S. The Integrative Action of the Nervous System ; Yale University Press: New Haven, CT, USA, 1906. [ Google Scholar ]
  • Zeng, H.; Sanes, J.R. Neuronal cell-type classification: Challenges, opportunities and the path forward. Nat. Rev. Neurosci. 2017 , 18 , 530–546. [ Google Scholar ] [ CrossRef ]
  • Dale, H.H. Chemical Transmission of the Effects of Nerve Impulses. BMJ 1934 , 1 , 835–841. [ Google Scholar ] [ CrossRef ]
  • Dale, H. Pharmacology and Nerve-Endings. Proc. R. Soc. Med. 1935 , 28 , 319–332. [ Google Scholar ]
  • Eccles, J.C. The Physiology of Nerve Cells ; John Hopkins Press: Baltimore, MD, USA, 1957. [ Google Scholar ]
  • Strata, P.; Harvey, R. Dale’s principle. Brain Res. Bull. 1999 , 50 , 349–350. [ Google Scholar ] [ CrossRef ]
  • Palay, S.L.; Wissig, S.L. Secretory granules and Nissl substance in fresh supraoptic neurones of the rabbit. Anat. Rec. 1953 , 116 , 301–313. [ Google Scholar ] [ CrossRef ]
  • Palade, G.E.; Palay, S.L. Electron microscope observations of interneuronal and neuromuscular synapses. Anat. Rec. 1954 , 118 , 335–336. [ Google Scholar ]
  • Palay, S.L. Synapses in the central nervous system. J. Biophys. Biochem. Cytol. 1956 , 2 , 193–202. [ Google Scholar ] [ CrossRef ]
  • Falck, B.; Hillarp, N.-Å.; Thieme, G.; Torp, A. Fluorescence of catechol amines and related compounds condensed with formaldehyde. J. Histochem. Cytochem. 1962 , 10 , 348–354. [ Google Scholar ] [ CrossRef ]
  • Dahlstrom, A.; Fuxe, K. Localization of monoamines in the lower brain stem. Experientia 1964 , 20 , 398–399. [ Google Scholar ] [ CrossRef ]
  • Anden, N.E.; Carlsson, A.; Dahlstroem, A.; Fuxe, K.; Hillrap, N.A.; Larsson, K. Demonstration and mapping out of nigro-neostriatal dopamine neurons. Life Sci. 1964 , 3 , 523–530. [ Google Scholar ] [ CrossRef ]
  • Fuxe, K. Evidence for the existence of monoamine neurons in the central nervous system: IV. Distribution of monoamine nerve terminals in the central nervous system. Acta Physiol. Scand. 1965 , 247 , 37–45. [ Google Scholar ]
  • Fuxe, K. Evidence for the existence of monoamine neurons in the central nervous system: 3. The monoamine nerve terminal. Z. Zellforsch. Mikrosk. Anat. 1965 , 65 , 573–596. [ Google Scholar ] [ CrossRef ]
  • Dahlström, A.; Fuxe, K. Evidence for the existence of monoamine neurons in the central nervous system: II. Experimentally induced changes in the intraneuronal amine levels of bulbospinal neuron systems. Acta Physiol. Scand. Suppl. 1965 , 247 , 1–36. [ Google Scholar ]
  • Benfenati, F.; Cimino, M.; Agnati, L.F.; Fuxe, K. Quantitative autoradiography of central neurotransmitter receptors: Methodological and statistical aspects with special reference to computer-assisted image analysis. Acta Physiol. Scand. 1986 , 128 , 129–146. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Agnati, L.F.; Fuxe, K.; Benfenati, F.; Zini, I.; Zoli, M.; Fabbri, L.; Härfstrand, A. Computer assisted morphometry and microdensitometry of transmitter identified neurons with special reference to the mesostriatal dopamine pathway—I. Methodological aspects. Acta Physiol. Scand. Suppl. 1984 , 52 , 5–36. [ Google Scholar ]
  • Zoli, M.; Zini, I.; Agnati, L.F.; Guidolin, D.; Ferraguti, F.; Fuxe, K. Aspects of neural plasticity in the central nervous system—I. Computer-assisted image analysis methods. Neurochem. Int. 1990 , 16 , 383–418. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Pert, C.B.; Kuhar, M.J.; Snyder, S.H. Autoradiograhic localization of the opiate receptor in rat brain. Life Sci 1975 , 16 , 1849–1853. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Hökfelt, T.; Johansson, O.; Goldstein, M. Chemical anatomy of the brain. Science 1984 , 225 , 1326–1334. [ Google Scholar ] [ CrossRef ]
  • Yelnik, J.; Bardinet, E.; Dormont, D.; Malandain, G.; Ourselin, S.; Tandé, D.; Karachi, C.; Ayache, N.; Cornu, P.; Agid, Y. A three-dimensional, histological and deformable atlas of human basal ganglia. I. Atlas construction based on immunohistochemical and MRI data. NeuroImage 2007 , 34 , 618–638. [ Google Scholar ] [ CrossRef ]
  • Hökfelt, T.; Elfvin, L.G.; Elde, R.; Schultzberg, M.; Goldstein, M.; Luft, R. Occurrence of somatostatin-like immunoreactivity in some peripheral sympathetic noradrenergic neurons. Proc. Natl. Acad. Sci. USA 1977 , 74 , 3587–3591. [ Google Scholar ] [ CrossRef ]
  • Hökfelt, T.; Holets, V.R.; Staines, W.; Meister, B.; Melander, T.; Schalling, M.; Schultzberg, M.; Freedman, J.; Björklund, H.; Olson, L.; et al. Coexistence of neuronal messengers—An overview. Prog. Brain Res. 1986 , 68 , 33–70. [ Google Scholar ] [ CrossRef ]
  • Furness, J.B.; Morris, J.L.; Gibbins, I.L.; Costa, M. Chemical coding of neurons and plurichemical transmission. Annu. Rev. Pharmacol. Toxicol. 1989 , 29 , 289–306. [ Google Scholar ] [ CrossRef ]
  • Burnstock, G. Cotransmission. Curr. Opin. Pharmacol. 2004 , 4 , 47–52. [ Google Scholar ] [ CrossRef ]
  • Svensson, E.; Apergis-Schoute, J.; Burnstock, G.; Nusbaum, M.P.; Parker, D.; Schiöth, H.B. General Principles of Neuronal Co-transmission: Insights from Multiple Model Systems. Front. Neural Circuits 2019 , 12 , 117. [ Google Scholar ] [ CrossRef ]
  • Hökfelt, T. Neuropeptides in perspective: The last ten years. Neuron 1991 , 7 , 867–879. [ Google Scholar ] [ CrossRef ]
  • Lundberg, J.M. Pharmacology of cotransmission in the autonomic nervous system: Integrative aspects on amines, neuropeptides, adenosine triphosphate, amino acids and nitric oxide. Pharmacol. Rev. 1996 , 48 , 113–178. [ Google Scholar ]
  • Hökfelt, T.; Broberger, C.; Xu, Z.D.; Sergeyev, V.; Ubink, R.; Diez, M. Neuropeptides—An overview. Neuropharmacology 2000 , 39 , 1337–1356. [ Google Scholar ] [ CrossRef ]
  • Cifuentes, F.; Morales, M.A. Functional Implications of Neurotransmitter Segregation. Front. Neural Circuits 2021 , 15 , 738516. [ Google Scholar ] [ CrossRef ]
  • Kandel, E.R. Cellular Basis of Behaviour ; W.H. Freeman: San Francisco, CA, USA, 1977. [ Google Scholar ]
  • Kandel, E.; Koester, J.D.; Mack, S.H.; Siegelbaum, S.A. Principles of Neuroscience , 6th ed.; McGraw Hill: New York, NY, USA, 2021. [ Google Scholar ]
  • Agnati, L.F.; Fuxe, K.; Zoli, M.; Pich, E.M.; Benfenati, F.; Zini, I.; Goldstein, M. Aspects on the information handling by the central nervous system: Focus on cotransmission in the aged rat brain. Prog. Brain Res. 1986 , 68 , 291–301. [ Google Scholar ] [ CrossRef ]
  • Kim, S.; Sabatini, B.L. Analytical approaches to examine gamma-aminobutyric acid and glutamate vesicular co-packaging. Front. Synaptic Neurosci. 2023 , 14 , 1076616. [ Google Scholar ] [ CrossRef ]
  • Hnasko, T.S.; Edwards, R.H. Neurotransmitter corelease: Mechanism and physiological role. Annu. Rev. Physiol. 2012 , 74 , 225–243. [ Google Scholar ] [ CrossRef ]
  • Tritsch, N.X.; Granger, A.J.; Sabatini, B.L. Mechanisms and fuction of GABA co-release. Nat. Rev. Neurosci. 2016 , 17 , 139–145. [ Google Scholar ] [ CrossRef ]
  • Saunders, A.; Oldenburg, I.A.; Berezovskii, V.K.; Johnson, C.A.; Kingery, N.D.; Elliott, H.L.; Xie, T.; Gerfen, C.R.; Sabatini, B.L. A direct GABAergic output from the basal ganglia to frontal cortex. Nature 2015 , 21 , 85–89. [ Google Scholar ] [ CrossRef ]
  • Changeaux, J.P.; Christopoulos, A. Allosteric modulation as a unifying mechanism for receptor function and regulation. Diabetes Obes. Metab. 2017 , 19 , 4–21. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Virchow, R. Ueber das granulirte Ansehen der Wandungen der Gehirnventrikel. In Gesammelte Abhandlungen zur Wissenschaftlichen Medicin ; Virchow, R., Ed.; Meidinger Sohn & Comp.: Frankfurt, Germany, 1856; pp. 885–891. [ Google Scholar ]
  • Virchow, R. Die Cellularpathologie in Ihrer Begründung auf Physiologische und Pathologische Gewebelehre ; Hirschwald: Berlin, Germany, 1858. [ Google Scholar ]
  • Lugaro, E. Sulle funzioni della nevroglia. Riv. Pat. Nerv. Ment. 1907 , 12 , 225–233. [ Google Scholar ]
  • Verkhratsky, A.; Nedergaard, M. Physiology of astroglia. Physiol. Rev. 2018 , 98 , 239–389. [ Google Scholar ] [ CrossRef ]
  • Sofroniew, M.V.; Vinters, H.V. Astrocytes: Biology and Pathology. Acta Neuropathol. 2010 , 119 , 7–35. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Parpura, V.; Heneka, M.T.; Montana, V.; Oliet, S.H.R.; Schousboe, A.; Haydon, P.G.; Stout, R.F., Jr.; Spray, D.C.; Reichenbach, A.; Pannicke, T.; et al. Glial cells in (patho)physiology. J. Neurochem. 2012 , 121 , 4–27. [ Google Scholar ] [ CrossRef ]
  • Guidolin, D.; Tortorella, C.; Marcoli, M.; Cervetto, C.; Maura, G.; Agnati, L.F. Receptor-Receptor Interactions and Glial Cell Functions with a Special Focus on G Protein-Coupled Receptors. Int. J. Mol. Sci. 2021 , 22 , 8656. [ Google Scholar ] [ CrossRef ]
  • Chvátal, A.; Syková, E. Glial influence on neuronal signaling. Prog. Brain Res. 2000 , 125 , 199–216. [ Google Scholar ] [ CrossRef ]
  • Färber, K.; Kettenmann, H. Physiology of microglial cells. Brain Res. Rev. 2005 , 48 , 133–143. [ Google Scholar ] [ CrossRef ]
  • Agnati, L.F.; Marcoli, M.; Maura, G.; Woods, A.; Guidolin, D. The brain as a “hyper-network”: The key role of neural networks as main producers of the integrated brain actions especially via the “broadcasted” neuroconnectomics. J. Neural Transm. 2018 , 125 , 883–897. [ Google Scholar ] [ CrossRef ]
  • Agnati, L.F.; Guidolin, D.; Maura, G.; Marcoli, M.; Leo, G.; Carone, C.; De Caro, R.; Genedani, S.; Borroto-Escuela, D.O.; Fuxe, K. Information handling by the brain: Proposal of a new “paradigm” involving the roamer type of volume transmission and the tunneling nanotube type of wiring transmission. J. Neural Transm. 2014 , 121 , 1431–1449. [ Google Scholar ] [ CrossRef ]
  • Fields, R.D.; Stevens-Graham, B. New insights into neuron-glia communication. Science 2002 , 298 , 556–562. [ Google Scholar ] [ CrossRef ]
  • Kofuji, P.; Araque, A. G-Protein-Coupled Receptors in Astrocyte-Neuron Communication. Neuroscience 2021 , 456 , 71–84. [ Google Scholar ] [ CrossRef ]
  • Sherwood, C.C.; Stimpson, C.D.; Raghanti, M.A.; Wildman, D.E.; Uddin, M.; Grossman, L.I.; Goodman, M.; Redmond, J.C.; Bonar, C.J.; Erwin, J.M.; et al. Evolution of increased glia-neuron ratios in the human frontal cortex. Proc. Natl. Acad. Sci. USA 2006 , 103 , 13606–13611. [ Google Scholar ] [ CrossRef ]
  • Von Bartheld, C.S.; Bahney, J.; Herculano-Houzel, S. The search for true numbers of neurons and glial cells in the human brain: A review of 150 years of cell counting. J. Comp. Neurol. 2016 , 524 , 3865–3895. [ Google Scholar ] [ CrossRef ]
  • Pelassa, S.; Guidolin, D.; Venturini, A.; Averna, M.; Frumento, G.; Campanini, L.; Bernardi, R.; Cortelli, P.; Buonaura, G.C.; Maura, G.; et al. A2A-D2 Heteromers on Striatal Astrocytes: Biochemical and Biophysical Evidence. Int. J. Mol. Sci. 2019 , 20 , 2457. [ Google Scholar ] [ CrossRef ]
  • Guidolin, D.; Marcoli, M.; Tortorella, C.; Maura, G.; Agnati, L.F. Adenosine A2A-Dopamine D2 Receptor-Receptor Interaction in Neurons and Astrocytes: Evidence and Perspectives. Prog. Med. Biol. Transl. Sci. 2020 , 169 , 247–277. [ Google Scholar ]
  • Amato, S.; Averna, M.; Guidolin, D.; Pedrazzi, M.; Pelassa, S.; Capraro, M.; Passalacqua, M.; Bozzo, M.; Gatta, E.; Anderlini, D.; et al. Heterodimer of A2A and oxytocin receptors regulating glutamate release in adult striatal astrocytes. Int. J. Mol. Sci. 2022 , 23 , 2326. [ Google Scholar ] [ CrossRef ]
  • Sporns, O. Cerebral cartography and connectomics. Philos. Trans. R. Soc. Lond. B Biol. Sci. 2015 , 370 , 20140173. [ Google Scholar ] [ CrossRef ]
  • Lord, L.-D.; Stevner, A.B.; Deco, G.; Kringelbach, M.L. Understanding principles of integration and segregation using whole-brain computational connectomics: Implications for neuropsychiatric disorders. Phil. Trans. R. Soc. A 2017 , 375 , 20160283. [ Google Scholar ] [ CrossRef ]
  • Harvey, R.J. Can computers think? Differences and similarities between computers and brains. Prog. Neurobiol. 1995 , 45 , 99–127. [ Google Scholar ] [ CrossRef ]
  • Shapshak, P. Artificial Intelligence and brain. Bioinformation 2018 , 14 , 38–41. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Singer, W. Differences Between Natural and Artificial Cognitive Systems. In Robotics, AI and Humanity ; von Braun, J., Archer, M.S., Reichberg, G.M., Sánchez Sorondo, M., Eds.; Springer: Cham, Switzerland, 2021; pp. 17–27. [ Google Scholar ] [ CrossRef ]
  • El Zaatari, S.; Marei, M.; Li, W.; Usman, Z. Cobot programming for collaborative industrial tasks: An overview. Robot. Auton. Syst. 2019 , 116 , 162–180. [ Google Scholar ] [ CrossRef ]
  • Guidolin, D.; Albertin, G.; Guescini, M.; Fuxe, K.; Agnati, L.F. Central nervous system and computation. Q. Rev. Biol. 2011 , 86 , 265–285. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Lee, D. Birth of Intelligence: From RNA to Artificial Intelligence ; Oxford University Press: Oxford, UK, 2020. [ Google Scholar ]
  • Agnati, L.F.; Anderlini, D.; Guidolin DMarcoli, M.; Maura, G. Man is a “Rope” Stretched Between Virosphere and Humanoid Robots: On the Urgent Need of an Ethical Code for Ecosystem Survival. Found. Sci. 2022 , 27 , 311–325. [ Google Scholar ] [ CrossRef ]
  • Guidolin, D.; Marcoli, M.; Maura, G.; Agnati, L.F. New dimensions of connectomics and network plasticity in the central nervous system. Rev. Neurosci. 2017 , 28 , 113–132. [ Google Scholar ] [ CrossRef ]
  • Agnati, L.F.; Fuxe, K.; Zoli, M.; Rondanini, C.; Ogren, S.O. New vistas on synaptic plasticity: The receptor mosaic hypothesis of the engram. Med. Biol. 1982 , 60 , 183–190. [ Google Scholar ]
  • Agnati, L.F.; Guidolin, D.; Leo, G.; Fuxe, K. A boolean network modelling of receptor mosaics relevance of topology and cooperativity. J. Neural. Transm. 2007 , 114 , 77–92. [ Google Scholar ] [ CrossRef ]
  • Dityatev, A.; Rusakov, D.A. Molecular signals of plasticity at the tetrapartite synapse. Curr. Opin. Neurobiol. 2011 , 21 , 353–359. [ Google Scholar ] [ CrossRef ]
  • Aggarwal, S.; Mortensen, O.V. Overview of Monoamine Transporters. Curr. Protoc. Pharmacol. 2017 , 79 , 12.16.1–12.16.17. [ Google Scholar ] [ CrossRef ]
  • Fuxe, K.; Dahlström, A.B.; Jonsson, G.; Marcellino, D.; Guescini, M.; Dam, M.; Manger, P.; Agnati, L. The discovery of central monoamine neurons gave volume transmission to the wired brain. Prog. Neurobiol. 2010 , 90 , 82–100. [ Google Scholar ] [ CrossRef ]
  • Zhou, Y.; Danbolt, N.C. Glutamate as a neurotransmitter in the healthy brain. J. Neural. Transm. 2014 , 121 , 799–817. [ Google Scholar ] [ CrossRef ]
  • Bernardinelli, Y.; Muller, D.; Nikonenko, I. Astrocyte-synapse structural plasticity. Neural Plast. 2014 , 2014 , 232105. [ Google Scholar ] [ CrossRef ]
  • Agnati, L.F.; Fuxe, K.; Zoli, M.; Ozini, I.; Toffano, G.; Ferraguti, F. A correlation analysis of the regional distribution of central enkephalin and beta endorphin immunoreactive terminals and of opiate receptors in adult and old male rats. Evidence for the existence of two main types of communication in the central nervous system: The volume transmission and the wiring transmission. Acta Physiol. Scand. 1986 , 128 , 201–207. [ Google Scholar ] [ CrossRef ]
  • Fuxe, K.; Agnati, L.F. Volume Transmission in the Brain, Novel Mechanisms for Neural Transmission ; Raven Press: New York, NY, USA, 1991; Volume 1. [ Google Scholar ]
  • Agnati, L.F.; Fuxe, K. Volume transmission as a key feature of information handling in the central nervous system: Possible new interpretative value of the Turing’s B-type machine. Progr. Brain Res. 2000 , 125 , 3–19. [ Google Scholar ] [ CrossRef ]
  • Marcoli, M.; Agnati, L.F.; Benedetti, F.; Genedani, S.; Guidolin, D.; Ferraro, L.; Maura, G.; Fuxe, K. On the role of the extracellular space on the holistic behavior of the brain. Rev. Neurosci. 2015 , 26 , 489–506. [ Google Scholar ] [ CrossRef ]
  • Kiss, J.P.; Vizi, E.S. Nitric oxide: A novel link between synaptic and nonsynaptic transmission. Trends Neurosci. 2001 , 24 , 211–215. [ Google Scholar ] [ CrossRef ]
  • Agnati, L.F.; Guidolin, D.; Guescini, M.; Genedani, S.; Fuxe, K. Understanding wiring and volume transmission. Brain Res. Rev. 2010 , 64 , 137–159. [ Google Scholar ] [ CrossRef ]
  • Taber, K.H.; Hurley, R.A. Volume transmission in the brain: Beyond the synapse. J. Neuropsychiatry Clin. Neurosci. 2014 , 26 , 4. [ Google Scholar ] [ CrossRef ]
  • Greitz, D. Cerebrospinal fluid circulation and associated intracranial dynamics. A radiologic investigation using MR imaging and radionuclide cisternography. Acta Radiol. Suppl. 1993 , 386 , 1–23. [ Google Scholar ]
  • Yablonskiy, D.A.; Ackerman, J.J.; Raichle, M.E. Coupling between changes in human brain temperature and oxidative metabolism during prolonged visual stimulation. Proc. Natl. Acad. Sci. USA 2000 , 97 , 7603–7608. [ Google Scholar ] [ CrossRef ]
  • Bullock, T.H. Signals and signs in the nervous system: The dynamic anatomy of electrical activity is probably information-rich. Proc. Natl. Acad. Sci. USA 1997 , 94 , 1–6. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Poo, M.M. Mobility and localization of proteins in excitable membranes. Annu. Rev. Neurosci. 1985 , 8 , 369–406. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Isakovic, J.; Dobbs-Dixon, I.; Chaudhuri, D.; Mitrecic, D. Modeling of inhomogeneous electromagnetic fields in the nervous system: A novel paradigm in understanding cell interactions, disease ethiology and therapy. Sci. Rep. 2018 , 8 , 12909. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Guescini, M.; Guidolin, D.; Vallorani, L.; Casadei, L.; Gioacchini, A.M.; Tibollo, P.; Battistelli, M.; Falcieri, E.; Battistin, L.; Agnati, L.F.; et al. C2C12 myoblasts release micro-vesicles containing mtDNA and proteins involved in signal transduction. Exp. Cell Res. 2010 , 316 , 1977–1984. [ Google Scholar ] [ CrossRef ]
  • Venturini, A.; Passalacqua, M.; Pelassa, S.; Pastorino, F.; Tedesco, M.; Cortese, K.; Gagliani, M.C.; Leo, G.; Maura, G.; Guidolin, D.; et al. Exosomes from Astrocyte Processes: Signaling to Neurons. Front. Pharmacol. 2019 , 10 , 1452. [ Google Scholar ] [ CrossRef ]
  • Guescini, M.; Leo, G.; Genedani, S.; Carone, C.; Pederzoli, F.; Ciruela, F.; Guidolin, D.; Stocchi, V.; Mantuano, M.; Borroto-Escuela, D.O.; et al. Microvesicle and tunneling nanotube mediated intercellular transfer of g-protein coupled receptors in cell cultures. Exp. Cell Res. 2012 , 318 , 603–613. [ Google Scholar ] [ CrossRef ]
  • Venero, J.L.; Vizuete, M.L.; Machado, A.; Cano, J. Aquaporins in the central nervous system. Prog. Neurobiol. 2001 , 63 , 321–366. [ Google Scholar ] [ CrossRef ]
  • Niermann, H.; Amiry-Moghaddam, M.; Holthoff, K.; Witte, O.W.; Ottersen, O.P. A novel role of vasopressin in the brain: Modulation of activity-dependent water flux in the neocortex. J. Neurosci. 2001 , 21 , 3045–3051. [ Google Scholar ] [ CrossRef ]
  • Sperelakis, N. Cell Physiology ; Academic Press: San Diego, CA, USA, 1997. [ Google Scholar ]
  • Jacob, F. Evolution and tinkering. Science 1977 , 196 , 1161–1166. [ Google Scholar ] [ CrossRef ]
  • Kandel, E.R.; Schwartz, J.H.; Jessel, T.M. Principles of Neural Science ; McGraw-Hill: New York, NY, USA, 2000. [ Google Scholar ]
  • McCulloch, W.S.; Pitts, W. A logical calculus of the ideas immanent in nervous activity. Bull. Math. Biophys. 1943 , 5 , 115–133. [ Google Scholar ] [ CrossRef ]
  • Woo, J.; Choi, K.; Kim, S.H.; Han, K.; Choi, M. Characterization of multiscale logic operations in the neural circuits. Front. Biosci. 2021 , 26 , 723–739. [ Google Scholar ] [ CrossRef ]
  • Agnati, L.F.; Fuxe, K.; Zini, I.; Lenzi, P.; Hökfelt, T. Aspects on receptor regulation and isoreceptor identification. Med. Biol. 1980 , 58 , 182–187. [ Google Scholar ]
  • Fuxe, K.; Agnati, L.F.; Benfenati, F.; Cimmino, M.; Algeri, S.; Hokfelt, T.; Mutt, V. Modulation by cholecystokinins of 3 H-spiroperidol binding in rat striatum: Evidence for increased affinity and reduction in the number of binding sites. Acta Physiol. Scand. 1981 , 113 , 567–569. [ Google Scholar ] [ CrossRef ]
  • Fuxe, K.; Agnati, L.F.; Benfenati, F.; Celani, M.; Zini, I.; Zoli, M.; Mutt, V. Evidence for the Existence of Receptor—Receptor Interactions in the Central Nervous System. Studies on the Regulation of Monoamine Receptors by Neuropeptides. J. Neural Transm. 1983 , 18 , 165–179. [ Google Scholar ] [ CrossRef ]
  • Agnati, L.F.; Guidolin, D.; Vilardaga, J.P.; Ciruela, F.; Fuxe, K. On the expanding terminology in the GPCR field: The meaning of receptor mosaics and receptor heteromers. J. Recept. Signal Transduct. Res. 2010 , 30 , 287–303. [ Google Scholar ] [ CrossRef ]
  • Agnati, L.F.; Fuxe, K.; Ferré, S. How receptor mosaics decode transmitter signals. Possible relevance of cooperativity. Trends Biochem. Sci. 2005 , 30 , 188–193. [ Google Scholar ] [ CrossRef ]
  • Agnati, L.F.; Celani, M.F.; Fuxe, K. Cholecystokinin peptides in vitro modulate the characteristics of the striatal 3H-Npropylnorapomorphine sites. Acta Physiol. Scand. 1983 , 118 , 79–81. [ Google Scholar ] [ CrossRef ]
  • Agnati, L.F.; Fuxe, K.; Benfenati, F.; Zini, I.; Hökfelt, T. On the functional role of coexistence of 5-HT and substance P in bulbospinal 5-HT neurons. Substance P reduces affinity and increases density of 3H-5-HT binding sites. Acta Physiol. Scand. 1983 , 117 , 299–301. [ Google Scholar ] [ CrossRef ]
  • Fuxe, K.; von Euler, G.; van der Ploeg, I.; Fredholm, B.B.; Agnati, L.F. Pertussis toxin treatment counteracts the cardiovascular effects of neuropeptide Y and clonidine in the awake unrestrained rat. Neurosci. Lett. 1989 , 101 , 337–341. [ Google Scholar ] [ CrossRef ]
  • Fuxe, K.; Agnati, L.F.; Harfstrand, A.; Zoli, M.; von Euler, G.; Grimaldi, R.; Merlo Pich, E.; Bjelke, B.; Eneroth, P.; Benfenati, F. On the role of neuropeptide Y in information handling in the central nervous system in normal and physio- pathological states. Focus on volume transmission and neuropeptide Y/alpha 2 receptor interactions. Ann. N. Y. Acad. Sci. 1990 , 579 , 28–67. [ Google Scholar ] [ CrossRef ]
  • Monod, J. Chance and Necessity: An Essay on the Natural Philosophy of Modern Biology ; William Collins Sons & Co. Ltd.: Glasgow, Scotland, 1979. [ Google Scholar ]
  • Fenton, A.W. Allostery: An illustrated definition for the ‘second secret of life’. Trends Biochem. Sci. 2008 , 33 , 420–425. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Agnati, L.F.; Tarakanov, A.O.; Guidolin, D. A simple mathematical model of cooperativity in receptor mosaics based on the “symmetry rule”. BioSystems 2005 , 80 , 165–173. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Dehner, C. Non-allosteric proteins. Why do proteins have quaternary structure? In Molecular Life Sciences. An Encyclopedic Reference ; Wells, R.D., Bond, J.S., Klinman, J., Siler Masters, B.S., Eds.; Springer: New York, NY, USA, 2018; pp. 812–818. [ Google Scholar ] [ CrossRef ]
  • Kenakin, T. Allosteric Theory: Taking Therapeutic Advantage of the Malleable Nature of GPCRs. Curr. Neuropharmacol. 2007 , 5 , 149–156. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Agnati, L.F.; Guidolin, D.; Leo, G.; Guescini, M.; Pizzi, M.; Stocchi, V.; Spano, P.F.; Ghidoni, R.; Ciruela, F.; Genedani, S.; et al. Possible new targets for GPCR modulation: Allosteric interactions, plasma membrane domains, intercellular transfer and epigenetic mechanisms. J. Recept. Signal Transduct. Res. 2011 , 31 , 315–331. [ Google Scholar ] [ CrossRef ]
  • Nussinov, R. The spatial structure of cell signaling systems. Phys. Biol. 2013 , 10 , 045004. [ Google Scholar ] [ CrossRef ]
  • Jordan, B.A.; Devi, L.A. G-protein-coupled receptor heterodimerization modulates receptor function. Nature 1999 , 399 , 697–700. [ Google Scholar ] [ CrossRef ]
  • George, S.R.; Fan, T.; Xie, Z.; Tse, R.; Tam, V.; Varghese, G.; O’Dowd, B.F. Oligomerization of mu- and delta-opioid receptors. Generation of novel functional properties. J. Biol. Chem. 2000 , 275 , 26128–26135. [ Google Scholar ] [ CrossRef ]
  • Rashid, A.J.; So, C.H.; Kong, M.M.; Furtak, T.; El-Ghundi, M.; Cheng, R.; O’Dowd, B.F.; George, S.R. D1-D2 dopamine receptor heterooligomers with unique pharmacology are coupled to rapid activation of Gq/11 in the striatum. Proc. Natl. Acad. Sci. USA 2007 , 104 , 654–659. [ Google Scholar ] [ CrossRef ]
  • Agnati, L.F.; Guidolin, D.; Cervetto, C.; Borroto-Escuela, D.O.; Fuxe, K. Role of iso-receptors in receptor-receptor interactions with a focus on dopamine iso-receptor complexes. Rev. Neurosci. 2016 , 27 , 1–25. [ Google Scholar ] [ CrossRef ]
  • Agnati, L.F.; Guidolin, D.; Fuxe, K. The brain as a system of nested but partially overlapping networks. Heuristic relevance of the model for brain physiology and pathology. J. Neural. Transm. 2007 , 114 , 3–19. [ Google Scholar ] [ CrossRef ]
  • Genedani, S.; Carone, C.; Guidolin, D.; Filaferro, M.; Marcellino, D.; Fuxe, K.; Agnati, L.F. Differential sensitivity of A2A and especially D2 receptor trafficking to cocaine compared with lipid rafts in cotransfected CHO cell lines. Novel actions of cocaine independent of the DA transporter. J. Mol. Neurosci. 2010 , 41 , 347–357. [ Google Scholar ] [ CrossRef ]
  • Hillion, J.; Canals, M.; Torvinen, M.; Casado, V.; Scott, R.; Terasmaa, A.; Hansson, A.; Watson, S.; Olah, M.E.; Mallol, J.; et al. Coaggregation, cointernalization, and codesensitization of adenosine A2A receptors and dopamine D2 receptors. J. Biol. Chem. 2002 , 277 , 18091–18097. [ Google Scholar ] [ CrossRef ]
  • Genedani, S.; Guidolin, D.; Leo, G.; Filaferro, M.; Torvinen, M.; Woods, A.S.; Fuxe, K.; Ferré, S.; Agnati, L.F. Computer-assisted image analysis of caveolin-1 involvement in the internalization process of adenosine A2A-dopamine D2 receptor heterodimers. J. Mol. Neurosci. 2005 , 26 , 177–184. [ Google Scholar ] [ CrossRef ]
  • Mores, K.L.; Cassell, R.J.; van Rijn, R.M. Arrestin recruitment and signaling by G protein-coupled receptor heteromers. Neuropharmacology 2019 , 152 , 15–21. [ Google Scholar ] [ CrossRef ]
  • Agnati, L.F.; Guidolin, D.; Carone, C.; Dam, M.; Genedani, S.; Fuxe, K. Understanding neuronal molecular networks builds on neuronal cellular network architecture. Brain Res. Rev. 2008 , 58 , 379–399. [ Google Scholar ] [ CrossRef ]
  • Agnati, L.F.; Fuxe, K.; Woods, A.; Genedani, S.; Guidolin, D. Theoretical considerations on the topological organization of receptor mosaics. Curr. Protein Pept. Sci. 2009 , 10 , 559–569. [ Google Scholar ] [ CrossRef ]
  • Agnati, L.F.; Guidolin, D.; Albertin, G.; Trivello, E.; Ciruela, F.; Genedani, S.; Tarakanov, A.; Fuxe, K. An integrated view on the role of receptor mosaics at perisynaptic level: Focus on adenosine A 2A , dopamine D 2 , cannabinoid CB 1 , and metabotropic glutamate mGlu 5 receptors. J. Recept. Signal Transduct. Res. 2010 , 30 , 355–369. [ Google Scholar ] [ CrossRef ]
  • Borroto-Escuela, D.O.; Brito, I.; Romero-Fernandez, W.; Di Palma, M.; Oflijan, J.; Skieterska, K.; Duchou, J.; Van Craenenbroeck, K.; Suárez-Boomgaard, D.; Rivera, A.; et al. The G protein-coupled receptor heterodimer network (GPCR-HetNet) and its hub components. Int. J. Mol. Sci. 2014 , 15 , 8570. [ Google Scholar ] [ CrossRef ]
  • Sriram, K.; Insel, P.A. G Protein-Coupled Receptors as Targets for Approved Drugs: How Many Targets and How Many Drugs? Mol. Pharmacol. 2018 , 93 , 251–258. [ Google Scholar ] [ CrossRef ]
  • Insel, P.A.; Sriram, K.; Gorr, M.W.; Wiley, S.Z.; Michkov, A.; Salmerón, C.; Chinn, A.M. GPCRomics: An Approach to Discover GPCR Drug Targets. Trends Pharmacol. Sci. 2019 , 40 , 378–387. [ Google Scholar ] [ CrossRef ]
  • Guo, S.; Zhao, T.; Yun, Y.; Xie, X. Recent progress in assays for GPCR drug discovery. Am. J. Physiol.-Cell Physiol. 2022 , 323 , C583–C594. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Maggio, R.; Fasciani, I.; Carli, M.; Petragnano, F.; Marampon, F.; Rossi, M.; Scarselli, M. Integration and Spatial Organization of Signaling by G Protein-Coupled Receptor Homo- and Heterodimers. Biomolecules 2021 , 11 , 1828. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Reiner, A.; Levitz, J. Glutamatergic Signaling in the Central Nervous System: Ionotropic and Metabotropic Receptors in Concert. Neuron 2018 , 98 , 1080–1098. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Scheefhals, N.; MacGillavry, H.D. Functional organization of postsynaptic glutamate receptors. Mol. Cell. Neurosci. 2018 , 91 , 82–94. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Borroto-Escuela, D.O.; Carlsson, J.; Ambrogini, P.; Narváez, M.; Wydra, K.; Tarakanov, A.O.; Li, X.; Millón, C.; Ferraro, L.; Cuppini, R.; et al. Understanding the Role of GPCR Heteroreceptor Complexes in Modulating the Brain Networks in Health and Disease. Front. Cell. Neurosci. 2017 , 11 , 37. [ Google Scholar ] [ CrossRef ]
  • Valle-León, M.; Callado, L.F.; Aso, E.; Cajiao-Manrique, M.M.; Sahlholm, K.; López-Cano, M.; Soler, C.; Altafaj, X.; Watanabe, M.; Ferré, S.; et al. Decreased striatal adenosine A 2A -dopamine D 2 receptor heteromerization in schizophrenia. Neuropsychopharmacology 2021 , 46 , 665–672. [ Google Scholar ] [ CrossRef ]
  • Valle-León, M.; Casajuana-Martin, N.; Del Torrent, C.L.; Argerich, J.; Gómez-Acero, L.; Sahlholm, K.; Ferré, S.; Pardo, L.; Ciruela, F. Unique effect of clozapine on adenosine A 2A -dopamine D 2 receptor heteromerization. Biomed. Pharmacother. 2023 , 160 , 114327. [ Google Scholar ] [ CrossRef ]
  • Fernández-Dueñas, V.; Gómez-Soler, M.; Valle-León, M.; Watanabe, M.; Ferrer, I.; Ciruela, F. Revealing Adenosine A 2A -Dopamine D 2 Receptor Heteromers in Parkinson’s Disease Post-Mortem Brain through a New AlphaScreen-Based Assay. Int. J. Mol. Sci. 2019 , 20 , 3600. [ Google Scholar ] [ CrossRef ]
  • Marcoli, M.; Agnati, L.F.; Franco, R.; Cortelli, P.; Anderlini, D.; Guidolin, D.; Cervetto, C.; Maura, G. Modulating brain integrative actions as a new perspective on pharmacological approaches to neuropsychiatric diseases. Front. Endocrinol. 2023 , 13 , 1038874. [ Google Scholar ] [ CrossRef ]
  • Li, K.; Li, J.; Zheng, J.; Qin, S. Reactive Astrocytes in Neurodegenerative Diseases. Aging Dis. 2019 , 10 , 664–675. [ Google Scholar ] [ CrossRef ]
  • Cragnolini, A.B.; Lampitella, G.; Virtuoso, A.; Viscovo, I.; Panetsos, F.; Papa, M.; Cirillo, G. Regional brain susceptibility to neurodegeneration: What is the role of glial cells? Neural Regen. Res. 2020 , 15 , 838–842. [ Google Scholar ] [ CrossRef ]
  • Diderot, D. Le Reve de d’Alembert ; Paris, France, 1830. [ Google Scholar ]
  • Guidolin, D.; Anderlini, D.; Maura, G.; Marcoli, M.; Cortelli, P.; Calandra-Buonaura, G.; Woods, A.S.; Agnati, L.F. A New Integrative Theory of Brain-Body-Ecosystem Medicine: From the Hippocratic Holistic View of Medicine to Our Modern Society. Int. J. Environ. Res. Public Health 2019 , 16 , 3136. [ Google Scholar ] [ CrossRef ]
  • Quigley, E.M.M. Microbiota-Brain-Gut Axis and Neurodegenerative Diseases. Curr. Neurol. Neurosci. Rep. 2017 , 17 , 94. [ Google Scholar ] [ CrossRef ]
  • Adak, A.; Khan, M.R. An insight into gut microbiota and its functionalities. Cell. Mol. Life. Sci. 2019 , 76 , 473–493. [ Google Scholar ] [ CrossRef ]
  • Shen, Y.; Xu, J.; Li, Z.; Huang, Y.; Yuan, Y.; Wang, J.; Zhang, M.; Hu, S.; Liang, Y. Analysis of gut microbiota diversity and auxiliary diagnosis as a biomarker in patients with schizophrenia: A cross-sectional study. Schizophr. Res. 2018 , 197 , 470–477. [ Google Scholar ] [ CrossRef ]
  • Calvani, R.; Picca, A.; Lo Monaco, M.R.; Landi, F.; Bernabei, R.; Marzetti, E. Of Microbes and Minds: A Narrative Review on the Second Brain Aging. Front. Med. 2018 , 5 , 53. [ Google Scholar ] [ CrossRef ]
  • Fülling, C.; Dinan, T.G.; Cryan, J.F. Gut Microbe to Brain Signaling: What Happens in Vagus…. Neuron 2019 , 101 , 998–1002. [ Google Scholar ] [ CrossRef ]
  • Fil, J.; Dalchau, N.; Chu, D. Programming Molecular Systems to Emulate a Learning Spiking Neuron. ACS Synth. Biol. 2022 , 11 , 2055–2069. [ Google Scholar ] [ CrossRef ]
  • Agnati, L.F.; Ferré, S.; Genedani, S.; Leo, G.; Guidolin, D.; Filaferro, M.; Carriba, P.; Casado, V.; Lluis, C.; Franco, R.; et al. Allosteric Modulation of Dopamine D 2 Receptors by Homocysteine. J. Proteome Res. 2006 , 5 , 3077–3083. [ Google Scholar ] [ CrossRef ]
  • Nussinov, R.; Tsai, C.-J.; Liu, J. Principles of Allosteric Interactions in Cell Signaling. J. Am. Chem. Soc. 2014 , 136 , 17692–17701. [ Google Scholar ] [ CrossRef ]
  • Schiffer, F. The physical nature of subjective experience and its interaction with the Brain. Med. Hypotheses 2019 , 125 , 57–69. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Friston, K.J. Waves of prediction. PLoS Biol. 2019 , 17 , e3000426. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Feynman, R.P. The Meaning of It All Thoughts of a Citizen Scientist ; Addison-Wesley: Boston, MA, USA, 1998. [ Google Scholar ]
  • Dickinson, E. J632 (1862). In The Complete Poems of Emily Dickinson ; Johnson, T.H., Ed.; Little Brown: Boston, MA, USA, 1890; p. 312. [ Google Scholar ]

Click here to enlarge figure

The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

Agnati, L.F.; Guidolin, D.; Cervetto, C.; Maura, G.; Marcoli, M. Brain Structure and Function: Insights from Chemical Neuroanatomy. Life 2023 , 13 , 940. https://doi.org/10.3390/life13040940

Agnati LF, Guidolin D, Cervetto C, Maura G, Marcoli M. Brain Structure and Function: Insights from Chemical Neuroanatomy. Life . 2023; 13(4):940. https://doi.org/10.3390/life13040940

Agnati, Luigi F., Diego Guidolin, Chiara Cervetto, Guido Maura, and Manuela Marcoli. 2023. "Brain Structure and Function: Insights from Chemical Neuroanatomy" Life 13, no. 4: 940. https://doi.org/10.3390/life13040940

Article Metrics

Article access statistics, further information, mdpi initiatives, follow mdpi.

MDPI

Subscribe to receive issue release notifications and newsletters from MDPI journals

Thank you for visiting nature.com. You are using a browser version with limited support for CSS. To obtain the best experience, we recommend you use a more up to date browser (or turn off compatibility mode in Internet Explorer). In the meantime, to ensure continued support, we are displaying the site without styles and JavaScript.

  • View all journals
  • Explore content
  • About the journal
  • Publish with us
  • Sign up for alerts
  • Perspective
  • Published: 08 November 2021

Anatomical structures, cell types and biomarkers of the Human Reference Atlas

  • Katy Börner   ORCID: orcid.org/0000-0002-3321-6137 1 ,
  • Sarah A. Teichmann   ORCID: orcid.org/0000-0002-6294-6366 2 ,
  • Ellen M. Quardokus 1 ,
  • James C. Gee 3 ,
  • Kristen Browne 4 ,
  • David Osumi-Sutherland 5 ,
  • Bruce W. Herr II   ORCID: orcid.org/0000-0002-6703-7647 1 ,
  • Andreas Bueckle   ORCID: orcid.org/0000-0002-8977-498X 1 ,
  • Hrishikesh Paul 1 ,
  • Muzlifah Haniffa   ORCID: orcid.org/0000-0002-3927-2084 6 ,
  • Laura Jardine 6 ,
  • Amy Bernard   ORCID: orcid.org/0000-0003-2540-1153 7 ,
  • Song-Lin Ding 8 ,
  • Jeremy A. Miller 8 ,
  • Shin Lin 9 ,
  • Marc K. Halushka 10 ,
  • Avinash Boppana 11 ,
  • Teri A. Longacre 12 ,
  • John Hickey 12 ,
  • Yiing Lin 13 ,
  • M. Todd Valerius   ORCID: orcid.org/0000-0001-8143-9231 14 ,
  • Yongqun He   ORCID: orcid.org/0000-0001-9189-9661 15 ,
  • Gloria Pryhuber 16 ,
  • Xin Sun 17 ,
  • Marda Jorgensen 18 ,
  • Andrea J. Radtke   ORCID: orcid.org/0000-0003-4379-8967 19 ,
  • Clive Wasserfall 18 ,
  • Fiona Ginty 20 ,
  • Jonhan Ho 21 ,
  • Joel Sunshine 22 ,
  • Rebecca T. Beuschel 19 ,
  • Maigan Brusko 18 ,
  • Sujin Lee 23 ,
  • Rajeev Malhotra   ORCID: orcid.org/0000-0003-0120-4630 14 , 23 ,
  • Sanjay Jain 24 , 25 &
  • Griffin Weber 26  

Nature Cell Biology volume  23 ,  pages 1117–1128 ( 2021 ) Cite this article

22k Accesses

48 Citations

30 Altmetric

Metrics details

  • Cellular imaging
  • Computational biology and bioinformatics
  • Molecular biology

The Human Reference Atlas (HRA) aims to map all of the cells of the human body to advance biomedical research and clinical practice. This Perspective presents collaborative work by members of 16 international consortia on two essential and interlinked parts of the HRA: (1) three-dimensional representations of anatomy that are linked to (2) tables that name and interlink major anatomical structures, cell types, plus biomarkers (ASCT+B). We discuss four examples that demonstrate the practical utility of the HRA.

Similar content being viewed by others

research paper on brain anatomy

Advances and prospects for the Human BioMolecular Atlas Program (HuBMAP)

research paper on brain anatomy

Specimen, biological structure, and spatial ontologies in support of a Human Reference Atlas

research paper on brain anatomy

Tissue registration and exploration user interfaces in support of a human reference atlas

With developments in massively parallel sequencing in bulk and at the single-cell level, researchers can now detect genomic features and genome expression with great precision 1 . Profiling single cells within tissues and organs enables researchers to map the distribution of cells and their developmental trajectories across organs and gives indications as to their functions. In 2021, there are several ongoing, ambitious efforts to map all of the cells in the human body and to create a digital reference atlas of the human body. The final atlas will encompass the three-dimensional (3D) organization of whole organs and thousands of anatomical structures, the interdependencies between trillions of cells, and the biomarkers that characterize and distinguish cell types. It will make the human body computable, supporting spatial and semantic queries run over 3D structures linked to their scientific terminology and existing ontologies. It will establish a benchmark reference that helps us to understand how the healthy human body works and what changes during ageing or disease.

A network of 16 consortia is contributing to the construction of the HRA based on studies of 30 organs (Fig. 1a ) with funding by the National Institutes of Health (NIH, blue) and strong support by the international Human Cell Atlas (HCA, red) 2 , 3 as well as expert input by reviewers from many different countries. The 16 consortia include the Allen Brain Atlas 4 , the Brain Research through Advancing Innovative Neurotechnologies Initiative—Cell Census Network Initiative 5 , the Chan Zuckerberg Initiative Seed Networks for HCA 2 , 3 , 6 , HCA awards by the EU’s Horizon 2020 program, the Genotype-Tissue Expression project 7 , the GenitoUrinary Developmental Molecular Anatomy Project 8 , Helmsley Charitable Trust: Gut Cell Atlas 2 , 3 , 6 , 9 , the Human Tumor Atlas Network 10 , the Human Biomolecular Atlas Program (HuBMAP) 11 , the Kidney Precision Medicine Project (KPMP) 12 , 13 , LungMAP 14 , HCA grants from the United Kingdom Research and Innovation Medical Research Council ( https://mrc.ukri.org ), (Re)building the Kidney 15 , Stimulating Peripheral Activity to Relieve Conditions 16 , The Cancer Genome Atlas 17 , 18 , 19 and Wellcome funding for HCA pilot projects 2 , 3 , 6 . In total, more than 2,000 experts from around the globe are working together to construct an open-source and free-to-use digital HRA using a wide variety of single or multimodal spatially resolved and bulk tissue assays. Imaging methods for anatomical structure segmentation include computed tomography, magnetic resonance imaging or optical coherence tomography (OCT) 20 . Spatially resolved single-cell methods detect metabolites or lipids using high-resolution nanospray desorption electrospray ionization mass spectrometry imaging 21 , proteins using co-detection by indexing 22 or tissue microarray-based immunohistochemistry 23 , simultaneous mRNA and chromatin accessibility using assay for transposase-accessible chromatin with high-throughput sequencing 24 , simultaneous protein and mRNA using cellular indexing of transcriptomes and epitopes by sequencing 25 , and mRNA using multiplexed error robust fluorescent hybridization (MERFISH) 26 , 27 , Slide-seq 28 , 29 , Nanostring’s GeoMX 30 or 10x Genomics Visium 31 .

figure 1

a , Alphabetical listing of 16 HRA construction efforts (left) linked to the 30 human organs that they study (right). The lungs are studied by ten consortia (orange links). This review focuses on ten organs (bold) plus vasculature. BICCN, Brain Research through Advancing Innovative Neurotechnologies Initiative—Cell Census Network Initiative; CZI, Chan Zuckerberg Initiative; H2020, Horizon 2020; GTEx, Genotype-Tissue Expression project; GUDMAP, GenitoUrinary Developmental Molecular Anatomy Project; HTAN, Human Tumor Atlas Network; MRC, Medical Research Council; RBK, (Re)building the Kidney; SPARC, Stimulating Peripheral Activity to Relieve Conditions; TCGA, The Cancer Genome Atlas. b , The 3D reference objects for major anatomical structures were jointly developed for 11 organs. c , An exemplary ASCT+B table showing anatomical structures (AS) and cell types (CT) and some biomarkers (B) for the glomerulus in the kidneys, annotated with the names of the three entity types (anatomical structures, cell types and biomarkers) and four relationship types (part_of, is_a, located_in and characterize). Note that the is_a relationship exists for cell types and biomarkers.

A major challenge in constructing the HRA is combining data generated by the different consortia without a common ‘language’ shared across them for describing and indexing the data in a spatially explicit and semantically consistent way. Rapid progress in single-cell technologies has led to an explosion of cell-type definitions. When researchers profile the expression of genes, proteins or other biomarkers, they can and do assign different cell types making it difficult, if not impossible, to compare results across studies—particularly across organ systems. Indeed, except for the brain 32 , no standards exist for the naming of anatomical structures, cell types and biomarkers. Furthermore, information on what cell types are commonly found in which anatomical structures and what biomarkers best characterize certain cell types is scattered across many ontologies (such as Uberon multi-species anatomy ontology 33 , the Foundational Model of Anatomy Ontology 34 , 35 , Cell Ontology (CL) 36 or the Human Gene Ontology Nomenclature Committee ( https://www.genenames.org/ )) and hundreds of publications about cells identified during development, disease and across multiple species (for example, the atlas efforts for brain 37 , heart 38 , lungs 14 and kidneys 13 ). Some critically important details (such as the morphology and distribution of microanatomical structures or the spatial layout of functionally interdependent cell types) are captured through hand-drawn figures—not digitally—without a shared 3D spatial reference system. The lack of a central, unified benchmark reference framework and language impedes progress in biomedical science as it is challenging or impossible to manage, compare, harmonize or use published data.

Towards a unified reference framework for mapping the human body

To address these issues, an NIH–HCA-organized meeting in March 2020 brought together leading experts to agree on major ontologies and associated 3D anatomical reference objects and to expand them as needed to capture the healthy human adult body. On the basis of ensuing discussions, more than 50 experts—including physicians, surgeons, anatomists, pathologists, experimentalists and representatives from the various consortia—have agreed on a reference framework to digitally represent relevant knowledge. The framework includes data structures, standard operating procedures and visual user interfaces that can be used by anatomists, pathologists, surgeons and other domain experts to digitize, integrate and analyse massive amounts of heterogeneous data. Experts also agreed on the major ontologies that will be used to create a ‘Rosetta Stone’ across existing anatomy, cell and biomarker ontologies. As a proof of concept, experts compiled inventories for 11 major organs (Fig. 1a,b ): bone marrow and blood plus pelvis reference organ 39 , 40 , 41 , 42 , 43 , 44 , 45 , 46 , 47 , 48 , 49 , 50 , 51 , 52 , brain 4 , 32 , 37 , 53 , 54 , heart 55 , 56 , large intestine 57 , 58 , 59 , 60 , 61 , 62 , 63 , 64 , 65 , 66 , 67 , 68 , kidneys 13 , 69 , 70 , 71 , 72 , 73 , 74 , 75 , 76 , 77 , lungs (refs. 78 , 79 , 80 , 81 , 82 , 83 , 84 , 85 , 86 and Sun, X. & Morrisey, E., manuscript in preparation), lymph nodes 87 , 88 , 89 , 90 , 91 , 92 , 93 , 94 , 95 , 96 , 97 , skin 98 , 99 , 100 , 101 , 102 , 103 , 104 , 105 , 106 , 107 , 108 , 109 , 110 , spleen 89 , 111 , 112 , 113 , 114 , 115 , 116 , 117 , 118 , 119 , 120 , 121 , 122 , thymus 123 , 124 , 125 , 126 , 127 , 128 , 129 , 130 , 131 , 132 , 133 , 134 and vasculature 78 , 135 , 136 , 137 , 138 , 139 , 140 , 141 . For each organ, the experts listed known anatomical structures, the cell types located in these structures and the biomarkers that are commonly used to characterize each cell type (such as gene and protein markers). The results are captured in ASCT+B tables (Fig. 1c ) that list and anatomical structures, cell types and biomarker entities, their relationships as well as references to supporting publications. As spatial position and context matter for cell function, the experts collaborated with medical designers to compile 3D, semantically annotated reference objects that cover the anatomically correct size and shape of major anatomical structures in a systematic and computable manner (see the final set for the initial 11 ASCT+B tables in Fig. 1b ). Together, the ASCT+B tables and associated 3D reference objects constitute the HRA.

An initial set of the ASCT+B tables and reference objects have been published ( Supplementary Information ). They capture data and knowledge that are mandatory for compiling a comprehensive HRA, and they are critically important for facilitating data exchange and collaboration among the 16 consortia and other efforts. They demonstrate how existing knowledge can be captured digitally and reorganized in support of a HRA. The tables and associated 3D reference objects provide an agreed-on framework for experimental data annotation across organs and scales (that is, from whole body to organs, tissues, cell types and biomarkers); they make it possible to compare and integrate data from different assay types (such as single-cell RNA-sequencing (scRNA-seq) 142 and MERFISH 143 data) for spatially equivalent tissue samples; they are a semantically and spatially explicit reference for healthy tissue and cell identity data that can then be compared against disease settings. Finally, the tables and associated reference organs can be used to evaluate progress on the semantic naming and definition of cell types and their anatomically accurate spatial characterization.

The ultimate goal is an HRA that correctly matches human diversity across populations. However, like the Human Genome Project 144 , 145 , the initial HRA covers a limited set of donors. For example, the Allen Human Reference Atlas—which is included in the HRA—was derived from one donor. It describes one half of a human brain that is intended to be mirrored to characterize a whole human brain. Most of the other reference organs in the current HRA are constructed from the Visible Human Project (VHP) male and female dataset made available by the National Library of Medicine 146 .

The presented ASCT+B tables and 3D reference objects are designed to be representative of organs and general enough to encompass all donors regardless of variation. All of the listed anatomical structures are expected to be present in a donor, except for sex-specific organs. There exists scientific evidence that the listed cell types are known to be located in the given anatomical structure, and that listed biomarkers are commonly used to identify a cell type. Future versions of the tables are in development and will make it possible to capture and compare data about variations in the size, location, shape and frequency of anatomical structures and cell types across donors. We next describe the data format, design and use of ASCT+B tables and the associated 3D reference objects. The initial HRA presents ten ASCT+B tables interlinked through a vasculature table together with a reference library of major anatomical structures. We discuss four examples that showcase the usage of the initial eleven-organ HRA for tissue registration and exploration, data integration, disease studies and measuring progress towards a more complete HRA. We conclude with a discussion of the next steps and an invitation to collaborate on the construction and usage of a reference atlas for healthy human adults.

ASCT+B tables

In 2019, the KPMP project published a first version of the ASCT+B tables to serve as a guide to annotate structures and cell types across multiple technologies in the kidneys 13 . The table data format was expanded and the process for constructing, reviewing and approving the ASCT+B tables was formalized (the key steps are described in Box 1 and the terminology is explained in Tables 1 and 2 ). Note that the HRA framework is in line with reproducibility best-practices and principles that make data findable, accessible, interoperable and reusable 147 , which are essential for the development of disease atlases, biomedical discovery and, ultimately, health improvements.

In September 2021, there exist 11 ASCT+B tables, version 1.0. These master tables are available for free online 148 . They capture 1,424 anatomical structures, 591 cell types and 1,867 biomarkers. The anatomical structures are linked by 2,543 ‘part_of’ relationships, 4,611 ‘located_in’ relationships between cell types and anatomical structures and 3,708 ‘characterize’ links between biomarkers and cells, supported by 293 unique scholarly publications and 506 web links ( Supplementary Information ). Like the first maps of our world, the first ASCT+B tables are imperfect and incomplete (see the ‘Limitations’ section). However, they digitize and standardize existing data and knowledge by clinicians, pathologists, anatomists and surgeons at the gross anatomical level; biologists, computer scientists and others at the single-cell level; and chemists, engineers and others at the biomarker level.

Box 1 Constructing the ASCT+B tables

At their core, the ASCT+B tables represent three entity types (anatomical structures, cell types and biomarkers; Fig. 1c ) and five relationship types (Fig. 1c and Table 2 ). Anatomical structures are connected through part_of relationships, creating a partonomy tree. In other words, the connections define whether anatomical structures may be shared parts within an organ or greater anatomical structure. Cell types are linked to other cell types through is_a relationships (for example, T cell is an immune cell, a cardiac cell is a muscle cell), defining the nature of the cell within a cellular lineage. Biomarkers can be of different types indicated by is_a (for example, they can be of type gene, protein, lipid, or metabolite); they are therefore defined in terms of their molecular nature. A bimodal network links cell types and anatomical structures on the basis of located_in relationships. Note that the same cell type might be located_in multiple anatomical structures, whereas a single anatomical structure might comprise multiple cell types. Biomarkers are linked to the cell types that they characterize through a second bimodal network. Note that one biomarker might be used to characterize multiple cell types, and multiple biomarkers might be required to uniquely characterize one cell type.

The ASCT+B v.1.0 table format makes it possible for human experts to represent the three entity types and five relationship types in a table. The initial set of 11 tables was authored using templated Google sheets. Experts filled in organ-specific tables by entering critical metadata (for example, authors, data and version number) in the top ten rows. Row 11 contains the header of the ASCT+B table, listing the anatomical structures, cell types, biomarkers and publication references from left to right. The table columns can be adjusted as needed (for example, anatomical structures partonomies might have only a few levels (5 for kidney), whereas others have many (19 for vasculature)). The v.1.0 format captures two biomarker types: gene markers (BG) and protein markers (BP); proteoforms, lipids and metabolites will be added in v1.1. Publication references document scientific evidence for the existence of the three entity types and their five interrelationships. The remaining rows, starting at row 12, contain the ASCT+B data—as many rows as there are unique cell types in the organ. Each unique anatomical structure, cell type and biomarker is represented by three columns. The first column lists the domain expert preferred name; whenever possible, this name should match the ontology name in the second column. The third column lists the unique, universally resolvable ontology ID, if available.

To ease table construction and ensure compliance with existing ontologies, a table of organ-specific, non-developmental human data captured in existing formalized ontologies was compiled by ontology experts and provided to the ASCT+B table authors. The ontologies used initially were Uberon 33 and Foundational Model of Anatomy for anatomical structures 34 , 35 , CL 36 for cell types and Human Gene Ontology Nomenclature Committee (HGNC) ( https://www.genenames.org/ ) for biomarkers. Data validation was performed by human experts and computationally by testing expert-curated relationships for validity in Uberon using Ubergraph 170 , a knowledge graph combining mutually referential Open Biological and Biomedical Ontology ontologies, including Cell Ontology and Uberon, and featuring precomputed classifications and relationships.

Box 2 Designing the 3D reference object library

To create male and female reference objects for the 11 initial organs, experts collaborated closely with medical designers to develop anatomically correct, vector-based objects that correctly represent human anatomy and are labelled using the ontology terms captured in the ASCT+B tables.

Data from the VHP male and female dataset, which was made available by the National Library of Medicine 146 , were used to model all of the 3D reference organs except for the brain, large intestine and lymph node. The brain uses the 141 anatomical structures of the ‘Allen Human Reference Atlas—3D, 2020’ representing one half of the human brain 4 ; these structures were mirrored to arrive at a whole human brain (as intended by the brain model authors) and resized to fit the visible human male and female bodies. A 3D model of the male large intestine was provided by A. Kaufman (Stony Brook University) modified to fit into the VHP male body, and used to guide the design of the female large intestine. The lymph node was created using mouse data and the clearing-enhanced 3D method developed by W. Li at the Laboratory of Immune System Biology, National Institute of Allergy and Infectious Diseases, NIH 171 . Although the size and cellular composition of mouse and human lymph nodes vary, the overall anatomy is well conserved between species.

All models were created using the medical image processing tool 3D Slicer 172 , 173 and modelling tools such as ZBrush 174 and Maya 175 . Files are provided in the Graphics Language Transmission Format (GLB) format 150 , which is a widely used standard file format for 3D scenes and models. GLB formats can be viewed in the free Babylon.js Sandbox ( https://sandbox.babylonjs.com/ ), making it possible for anyone to explore the 3D reference objects using a web browser, without downloading or installing new software.

Three-dimensional reference object library

The spatial location of cell types within anatomical structures matters, as do the number and types of cells within the same anatomical structure. A 3D reference object library was compiled (key steps are described in Box 2) to capture the size, shape, position and rotation of major anatomical structures in the organ-specific ASCT+B tables (Box 1 , Tables 1 and 2 , and Supplementary Information ).

In September 2021, there exist 26 organ objects that complement the 11 ASCT+B tables discussed above. Male and female versions exist for all organs, and left and right versions exist for kidneys and lymph nodes. All organs are properly positioned in the male and female bodies and are freely available online 149 . The resulting 3D reference object library captures a total of 1,185 unique 3D structures (for example, the left female kidney has 11 renal papillae; see the complete listing in the HuBMAP Consortium 3D Reference Object Library 149 ) with 557 unique Uberon terms, including the name of the organ. Files are provided in the GLB format 150 and are used in several user interfaces (see the ‘Tissue registration and exploration’ section).

A ‘crosswalk’ mapping file links the anatomical structure names listed in the ASCT+B tables and the terminology used in the 3D reference object library GLB files (see the ‘CCF 3D reference object library & crosswalk’ section in the Supplementary Information ). This interlinkage of ontology terms (for example, kidney or cortex of kidney; Fig. 1c ), and the 3D references objects with well-defined polygon meshes (Table 1 ) that describe the size and shape of these anatomical structures in 3D (Fig. 1b ) is critical for using 3D collision-detection algorithms for tissue registration and spatial search (see the ‘Tissue registration and exploration’ section).

How to use the HRA

The ASCT+B tables and associated 3D reference organ objects provide a starting point for the systematic construction of a HRA. They provide common nomenclature for major entities and relationships along with cross-references to existing ontologies and supporting literature. In the following sections, we provide four examples illustrating the value of the tables and associated 3D reference organs: (1) to support experimental tissue data registration and annotation across organs and scales (see the next section); (2) to compare and integrate data from different assay types (see the ‘Comparing cell states across different tissues and in disease’ section); (3) to compare healthy and disease data (see the ‘Understanding disease’ section); (4) and to evaluate progress toward the compilation of a comprehensive HRA (see the ‘Measuring progress’ section).

Tissue registration and exploration

Like any atlas, a HRA is a collection of maps that capture a multiscale 3D reality. Like other digital maps, it supports panning and zoom, from the whole body (macro scale; metres), to the organ level (meso scale; centimetres), to the level of functional tissue units (such as alveoli in the lungs, crypts in the colon and glomeruli in the kidneys; millimetres), down to the single-cell level (micro scale; micrometres). To be usable, the maps in an HRA must use the same index terms and a unifying topological coordinate system such that cells and anatomical structures in adjacent overlapping maps or at different zoom levels can be uniquely named and properly aligned.

The ASCT+B tables and 3D reference organs provide a framework for experimental data annotation and exploration across organs and scales—that is, from the entire body down to organs, tissues, cell types and biomarkers. For example, they are used to support the registration of new tissue data as well as spatial and semantic search, browsing and exploration of human tissue data. The HuBMAP CCF Registration User Interface 151 (RUI; Fig. 2a ) and CCF Exploration User Interface (EUI) 152 (Fig. 2b ) are available at the HuBMAP portal 153 . The code for both user interfaces is freely available at GitHub HuBMAP Consortium CCF User Interfaces 154 ( https://github.com/hubmapconsortium/ccf-ui ). Several consortia and single investigators have used the stand-alone version of the CCF RUI 151 to register human tissue samples. Resulting data have been added to the CCF EUI making it possible for data providers and others to explore these tissue datasets in the context of human anatomy, to search and filter for datasets that match certain criteria (such as a specific age, sex and assay type) and to access and download raw data.

figure 2

a , Registration and anatomical structure annotation of tissue data (blue block) in 3D through collision detection in the RUI. A user sizes, positions and rotates tissue blocks, and saves the results in the JavaScript Object Notation (JSON) format. b , Tissue data can be queried, filtered and explored through the EUI. RUI-registered tissue data (white blocks in the spleen and kidneys) can be explored semantically using the anatomical structure partonomy on the left and spatially using the anatomy browser in the middle; a filter at the top right supports subsetting by sex, age, tissue provider and so on. Clicking on a tissue sample on the right links to the Vitessce image viewer 169 .

To make it easy for anyone to explore or contribute HRA data, we developed diverse learning modules as part of the free Visible Human Massive Open Online Course ( https://expand.iu.edu/browse/sice/cns/courses/hubmap-visible-human-mooc ). This course describes the compilation and coverage of HRA data, demonstrates new single-cell analysis and mapping techniques, and introduces the diverse HRA user interfaces. Delivered entirely online, all coursework can be completed asynchronously to fit busy schedules.

Comparing cell states across different tissues and in disease

The ASCT+B framework provides a ‘look-up table’ from the 3D reference models to the unique ontology name and ID for anatomical structures and their cell-type composition across organs that are present in the ASCT+B tables. Researchers can use the tables to determine in what anatomical structure a cell type is commonly located. Specifically, the ASCT+B tables capture information on cells formed within and resident in a specific tissue (such as epithelia and stroma) as well as cells that migrate across tissues (such as immune cells) 155 . For example, immune cells originate primarily in the bone marrow in postnatal life. Adaptive lymphocytes subsequently differentiate and mature in lymphoid tissues such as the thymus and spleen before circulating to non-lymphoid tissues and lymph nodes (Fig. 3a ). Thus, these cells recur across the ASCT+B tables in both the lymphoid (bone marrow, thymus, spleen, lymph node) and non-lymphoid (brain, heart, kidney, lung, skin) tissue tables. Existing data support a more nuanced and tissue-specific, ontology-based assignment of blood and immune cells. For example, scRNA-seq enables deep phenotyping of haematopoietic stem cells (HSCs) and haematopoietic stem and progenitor cells (HSPCs) and their differentiated progenies across various tissues. When comparing fetal liver versus thymic cell states (Fig. 3a (right)), a small region of the HSPCs highlighted as lymphoid progenitors is shared across the two organs, indicating cells that have migrated from the liver to the thymus 52 , 125 . Data integration defines molecules (such as chemokine receptors) that determine tissue residency versus migratory properties. These biomarkers in turn define tissue-resident versus migratory cell states, which can be added to the ASCT+B tables to refine cellular ontology.

figure 3

a , HSCs (CL:0000037) migrate from the liver (UBERON:0002107) to the thymus (UBERON:0002370) during embryonic and fetal development. The transcriptomic identity of these HSCs changes throughout pregnancy. The differences in these anatomical structures are shown by the maroon versus blue shading of the HSCs on the left (embryo) and right (fetal) parts of the figure. The scRNA-seq data in the uniform manifold approximation and projection plots are from a published single-cell transcriptomic profile of the thymus across the human lifetime 125 . The top plot shows liver cells (blue) and thymus cells (orange) overlapping, which are labelled lymphoid progenitors in the bottom plot. The other cell populations shown include HSCs; double negative T cells (DNT); megakaryocytes (MKs); megakaryocyte/erythrocyte/mast cell progenitors (MEMPs); pro-B cells; neutrophil–myeloid progenitors (NMPs); mast cells; and erythrocytes (Ery). b , The nature of HSC subsets in the adult blood shifts in health versus COVID-19. HSCs in the blood of patients with COVID-19 (top left) show a megakaryocyte priming bias compared with healthy cells (top right). This is quantified in the histogram from the human thymus single-cell atlas of relative HSPC contributions for different donor/patient cohorts 156 .

A well-annotated healthy HRA can then be used to understand the molecular and cellular alterations in response to perturbations such as infection. For example, data from several single-cell multi-omics studies of patients’ blood can be combined to compute the cellular response during COVID-19 pathogenesis, including HSC progenitor states that emerge during disease 156 (Fig. 3b ).

Understanding disease

As a reference for healthy tissue, the ASCT+B tables can be used to identify changes in molecular states in normal ageing or disease. For example, the kidney master table links relevant anatomical structures, cell types and biomarkers to disease and other ontologies for increasing our understanding of disease states. The top significant and specific biomarkers in each reference cell/state cluster might differ during disease or in a cell undergoing repair, regeneration, or in a state of failed or maladaptive repair. Loss of expression or alteration in the cellular distribution of a specific biomarker may provide clues to the underlying disease. The Kidney Precision Medicine Project is working towards ASCT+B tables that characterize disease. Researchers aim to include biomarkers with important physiological roles in maintaining the cellular architecture or function and biomarkers that reveal shifts in cell types that are associated with acute and chronic diseases 157 . Changes in biomarkers in healthy and injured cells provide information about the underlying biological pathways that drive these shifts and therefore provide critical insights into pathogenic mechanisms. For example, the gene NPHS1 , which encodes nephrin, is one of the top markers of healthy podocytes and is essential for glomerular function. Mutations in NPHS1 may be found in patients with proteinuria 158 . The kidney ASCT+B table records that the gene biomarker NPHS1 (Fig. 1c (bottom right)) is expressed in the podocytes of the kidney. Ontology suggests injury to podocytes and glomerular function may cause proteinuria (Fig. 4 ). Ontology IDs provided for anatomical structures, cell types and biomarkers facilitate linkages to clinicopathological knowledge and help to provide broader insights into disease 159 . For example, the ASCT+B kidney master table and single-nucleus RNA-seq atlas data 70 have been used to characterize diabetic nephropathy disease states by distinguishing the healthy interstitium from a diabetic one 160 . Note that some of the existing data are not at the single-cell level; in these cases, regional data (such as data bounded by tissue blocks registered within reference organs with known anatomical structures, cell types and biomarkers (see the RUI and EUI discussion above)) can be compared to the kidney master table. In summary, ASCT+B tables interlinked with existing ontologies provide a foundation for new data analysis and the functional study of diseases.

figure 4

Processes, anatomical structures, cell types, biomarkers and disease relevant for understanding proteinuria, in kidney. The example shows how the ASCT+B tables help link clinicopathology information and ontology data.

Measuring progress

The ASCT+B tables provide objective measures for tracking progress towards an accurate and complete HRA. In particular, if a scholarly publication includes a new ASCT+B table, that table can be compared with existing master tables and the number and type of identical (confirmatory) and different (new) anatomical structures, cell types and biomarkers, as well as their relationships, can be determined. The value of a new data release for reference atlas design can be evaluated in terms of the number and type of new anatomical structures, cell types, biomarkers and their relationships that it contributes. The ASCT+B Reporter 161 supports the visual exploration and comparison of ASCT+B tables. Table authors and reviewers can use this online tool to upload new tables, examine them visually and compare them with existing master tables. Power analysis methods can be run to assess the coverage and completeness of cell states and/or types and decide what tissues and cells should be sampled next 139 in support of a data-driven experimental design.

As the number of published ASCT+B tables grows, estimates can be run to determine the likely accuracy of anatomical structures, cell types, biomarkers and relationships. Entities and relationships based on high-quality data or multiple types of scholarly evidence are more likely to be correct compared with those with limited or no evidence. Incomplete data can be easily identified and flagged (for example, anatomical structures with no linkages to cell types and cell types with no biomarkers indicate missing data). Given that the tables disclose who contributes the data and who authors relevant publications, experts on anatomical structures, cell types, biomarkers and their relationships can be identified and invited to further improve the tables.

Limitations

The current format of the ASCT+B tables and 3D reference organs makes them easy to author, review, validate and use across organs and domains of expertise. The ASCT+B Reporter tool 161 supports the authoring and review of ASCT+B tables but also the comparison of these data with new datasets. 3D reference objects can be freely explored in the Babylon.js sandbox in a web browser ( https://sandbox.babylonjs.com/ ). However, the simplicity of the current tables and organs makes it impossible to fully capture the complexity of the human body. Thus, for each organ-specific table, experts recorded the process that they used to construct the table, which often included simplifying the anatomy to fit within a strict partonomy, making decisions about which cell types and biomarkers had sufficient evidence to be included in the table, or ignoring normal dynamic changes that occur in the organ over time. For several organs, such as the brain, the biomarkers are preliminary and are expected to improve in coverage and robustness in the future.

Biases in sampling with respect to donor demographics (for example, from convenience samples as opposed to using sampling strategies that reflect global demographics), organs (for example, as based on availability of funding) or cell types (for example, due to differential viability or capture efficiency) can be determined and need to be proactively addressed to arrive at an atlas that truly captures healthy human adults. The definition of ‘healthy’ is expected to evolve; HRA metadata for datasets that were used in the construction of the atlas include information on sex, age and ethnicity, but also comorbidities, making it possible to include or exclude datasets when inclusion criteria change and to recompute the HRA as needed.

Diversity, inclusion, and global scientific equity are major goals for all of the consortia involved in this effort 162 . Authors for the initial set of tables are from the United States and United Kingdom, but many have roots, training and collaborative connections in other parts of the world. Reviewers do not solely come from the United States and United Kingdom. Aiming to overcome the impact of COVID-19 related travel restrictions on collaboration across disciplinary, institutional and cultural boundaries, we organized the HRA panel at the virtual Spatial Biology Europe meeting in April 2021. The panel featured presentations by experts from Asia, Australia and North America. Slides and recordings are available online 163 . In the future, we hope to engage experts from many more countries as well as students from diverse backgrounds.

ASCT+B tables in combination with the 3D reference object library provide a rigorous cross-organ framework for experimental data annotation and exploration from the levels of organs and tissues to the levels of cell types and biomarkers. The construction and validation of the tables are iterative. Initially, ontology and publication data, along with knowledge from organ experts, are codified and unified. Later, experimental datasets are compared with existing master tables to confirm the tables or add to them as needed to capture healthy human tissue data. In the near future, the cell type typology will be expanded from one level to multiple levels, also using data from the evolving set of Azimuth references 164 , 165 and in close collaboration with the Cell Ontology curation effort (see ref. 166 in this issue). This will make it possible to compare anatomical structure partonomy and cell-type typology datasets at different levels of resolution. New organs will be added to the 3D reference library and microanatomical structures, such as glomeruli in the kidneys, crypts in the large intestine and alveoli in the lungs will be included.

Efforts are underway to integrate cell-specific protein biomarkers in the ASCT+B tables with well-characterized antibodies for multiplexed antibody-based imaging; the ASCT+B reporter has already contributed to detect protein biomarkers in situ 167 , 168 . The goal of these efforts is to generate expertly curated, tissue-specific antibody panels that can be used across consortia in support of a HRA.

The number of anatomical structures, cell types and robust biomarkers will probably increase as new single-cell technologies and computational workflows are developed. Thus, the tables and associated reference objects are a living ‘snapshot’ of the status of the collective work towards an open, authoritative, computable HRA, against which experimentalists can calibrate their data and to which they can contribute. Future uses of the HRA might include cross-species comparisons or cross-species annotations, cross-tissue/organ comparisons, comparisons of healthy versus common or rare genetic variations, and usage in teaching—expanding widely used anatomy books 78 , 135 to the single-cell level.

It will take extensive effort and expertise to arrive at a consensus HRA and to develop methods and user interfaces that use it to advance research and improve human health. Experts interested in contributing to this international and interdisciplinary effort are invited to register at https://iu.co1.qualtrics.com/jfe/form/SV_bpaBhIr8XfdiNRH to receive regular updates and invites to meetings that aim to advance the construction of the HRA.

Chen, G., Ning, B. & Shi, T. Single-cell RNA-seq technologies and related computational data analysis. Front. Genet. https://doi.org/10.3389/fgene.2019.00317 (2019).

Rozenblatt-Rosen, O., Stubbington, M. J. T., Regev, A. & Teichmann, S. A. The human cell atlas: from vision to reality. Nature 574 , 187–192 (2017).

Google Scholar  

Regev, A. et al. The Human Cell Atlas. eLife 6 , e27041 (2017).

Article   PubMed   PubMed Central   Google Scholar  

Ding, S. L. et al. Comprehensive cellular-resolution atlas of the adult human brain. J. Comp. Neurol. 524 , 3127–3481 (2016).

Devor, A. et al. The challenge of connecting the dots in the B.R.A.I.N. Neuron 80 , 270–274 (2013).

Article   CAS   PubMed   Google Scholar  

Moghe, I., Loupy, A. & Solez, K. The human cell atlas project by the numbers: relationship to the Banff classification. Am. J. Transpl. 18 , 1830 (2018).

Article   Google Scholar  

Lonsdale, J. et al. The Genotype-Tissue Expression (GTEx) project. Nat. Genet. 45 , 580–585 (2013).

Article   CAS   Google Scholar  

McMahon, A. P. et al. GUDMAP: the genitourinary developmental molecular anatomy project. J. Am. Soc. Nephrol. 19 , 667–671 (2008).

Article   PubMed   Google Scholar  

Elmentaite, R. et al. Cells of the human intestinal tract mapped across space and time. Nature 597 , 250–255 (2021).

Article   CAS   PubMed   PubMed Central   Google Scholar  

Srivastava, S. et al. The making of a PreCancer Atlas: promises, challenges, and opportunities. Trends Cancer 4 , 523–536 (2018).

Snyder, M. P. et al. The human body at cellular resolution: the NIH Human Biomolecular Atlas Program. Nature 574 , 187–192 (2019).

Himmelstein, D. S. et al. Systematic integration of biomedical knowledge prioritizes drugs for repurposing. eLife 6 , e26726 (2017).

El-Achkar, T. M. et al. A multimodal and integrated approach to interrogate human kidney biopsies with rigor and reproducibility: guidelines from the Kidney Precision Medicine Project. Physiol. Genomics 53 , 1–11 (2021).

Ardini-Poleske, M. E. et al. LungMAP: the molecular atlas of lung development program. Am. J. Physiol. Lung Cell. Mol. Physiol. 313 , L733–L740 (2013).

Oxburgh, L. et al. (Re)building a kidney. J. Am. Soc. Nephrol. 28 , 1370–1378 (2017).

Stimulating Peripheral Activity to Relieve Conditions (SPARC) (NIH, 2020); https://commonfund.nih.gov/sparc

Heng, H. H. Q. Cancer genome sequencing the challenges ahead. Bioessays 29 , 783–794 (2007).

TGCA Research Network Comprehensive genomic characterization defines human glioblastoma genes and core pathways. Nature 455 , 1061–1068 (2008).

Freire, P. et al. Exploratory analysis of the copy number alterations in glioblastoma multiforme. PLoS ONE 3 , e4076 (2008).

Aumann, S., Donner, S., Fischer, J. & Muller, F. in High Resolution Imaging in Microscopy and Ophthalmology: New Frontiers in Biomedical Optics (ed. Bille, J. F.) 59–85. Springer (2019).

Yin, R., Burnum-Johnson, K. E., Sun, X., Dey, S. K. & Laskin, J. High spatial resolution imaging of biological tissues using nanospray desorption electrospray ionization mass spectrometry. Nat. Protoc. 14 , 3445–3470 (2019).

Goltsev, Y. et al. Deep profiling of mouse splenic architecture with CODEX multiplexed imaging. Cell 174 , 968–981 (2018).

Uhlén, M. et al. Tissue-based map of the human proteome. Science 347 , 1260419 (2015).

Buenrostro, J. D. et al. Single-cell chromatin accessibility reveals principles of regulatory variation. Nature 523 , 486–490 (2015).

Stoeckius, M. et al. Simultaneous epitope and transcriptome measurement in single cells. Nat. Methods 14 , 865–868 (2017).

Chen, K. H., Boettiger, A. N., Moffitt, J. R., Wang, S. & Zhuang, X. Spatially resolved, highly multiplexed RNA profiling in single cells. Science 348 , aaa6090 (2015).

Moffitt, J. R. et al. High-throughput single-cell gene-expression profiling with multiplexed error-robust fluorescence in situ hybridization. Proc. Natl Acad. Sci. USA 113 , 11046–11051 (2016).

Rodriques, S. G. et al. Slide-seq: a scalable technology for measuring genome-wide expression at high spatial resolution. Science 363 , 1463–1467 (2019).

Asp, M., Bergenstrahle, J. & Lundeberg, J. Spatially resolved transcriptomes-next generation tools for tissue exploration. Bioessays 42 , e1900221 (2020).

Zollinger, D. R., Lingle, S. E., Sorg, K., Beechem, J. M. & Merritt, C. R. GeoMx RNA Assay: high multiplex, digital, spatial analysis of RNA in FFPE tissue. Methods Mol. Biol. 2148 , 331–345 (2020).

Visium Spatial Gene Expression (10x Genomics, 2021); https://www.10xgenomics.com/products/spatial-gene-expression

Miller, J. A. et al. Common cell type nomenclature for the mammalian brain. eLife 9 , e59928 (2020).

Mungall, C. J., Torniai, C., Gkoutos, G. V., Lewis, S. E. & Haendel, M. A. Uberon, an integrative multi-species anatomy ontology. Genome Biol. 13 , R5 (2012).

Golbreich, C., Grosjean, J. & Darmoni, S. J. The foundational model of anatomy in OWL 2 and its use. Artif. Intell. Med. 57 , 119–132 (2013).

Rosse, C. & Mejino, J. L. V. A reference ontology for biomedical informatics: the Foundational Model of Anatomy. J. Biomed. Inform. 36 , 478–500 (2003).

Meehan, T. F. et al. Logical development of the Cell Ontology. BMC Bioinform. 12 , 6 (2011).

Ding, S. L. et al. Allen Human Reference Atlas—3D, 2020 (2021); http://download.alleninstitute.org/informatics-archive/allen_human_reference_atlas_3d_2020/version_2021

Fonseca, C. G. et al. The Cardiac Atlas Project: an imaging database for computational modeling and statistical atlases of the heart. Bioinformatics 27 , 2288–2295 (2011).

Géron, A., Werner, J., Wattiez, R., Lebaron, P. & Mattallana-Surget, S. Deciphering the functioning of microbial communities: shedding light on the critical steps in metaproteomics. Front. Microbiol. 10 , 2395 (2019).

Manz, M. G., Miyamoto, T., Akashi, K. & Weissman, I. L. Prospective isolation of human clonogenic common myeloid progenitors. Proc. Natl Acad. Sci. USA 99 , 11872–11877 (2002).

Fajtova, M., Kovarikova, A., Svec, P., Kankuri, E. & Sedlak, J. Immunophenotypic profile of nucleated erythroid progenitors during maturation in regenerating bone marrow. Leuk. Lymphoma 54 , 2523–2530 (2013).

Kawamura, S. et al. Identification of a human clonogenic progenitor with strict monocyte differentiation potential: a counterpart of mouse cMoPs. Immunity 54 , 2523–2530 (2017).

Mousset, C. M. et al. Comprehensive phenotyping of T cells using flow cytometry. Cytom. A 95 , 647–654 (2019).

Mello, F. V. et al. Maturation-associated gene expression profiles along normal human bone marrow monopoiesis. Br. J. Haematol. 176 , 464–474 (2017).

Tomer, A. Human marrow megakaryocyte differentiation: multiparameter correlative analysis identifies von Willebrand factor as a sensitive and distinctive marker for early (2N and 4N) megakaryocytes. Blood 104 , 2722–2727 (2004).

Doulatov, S. et al. Revised map of the human progenitor hierarchy shows the origin of macrophages and dendritic cells in early lymphoid development. Nat. Immunol. 11 , 585–593 (2010).

Elghetany, M. T., Ge, Y., Patel, J., Martinez, J. & Uhrova, H. Flow cytometric study of neutrophilic granulopoiesis in normal bone marrow using an expanded panel of antibodies: correlation with morphologic assessments. J. Clin. Lab. Anal. 18 , 36–41 (2004).

Szabo, P. A. et al. Single-cell transcriptomics of human T cells reveals tissue and activation signatures in health and disease. Nat. Commun. 10 , 4706 (2019).

Kaminski, D. A., Wei, C., Qian, Y., Rosenberg, A. F. & Sanz, I. Advances in human B cell phenotypic profiling. Front. Immunol. 3 , 302 (2012).

Hay, S. B., Ferchen, K., Chetal, K., Grimes, H. L. & Salomonis, N. The Human Cell Atlas bone marrow single-cell interactive web portal. Exp. Hematol. 68 , P51–P61 (2018).

Clavarino, G. et al. Novel strategy for phenotypic characterization of human B lymphocytes from precursors to effector cells by flow cytometry. PLoS ONE 11 , e0162209 (2016).

Popescu, D. M. et al. Decoding human fetal liver haematopoiesis. Nature 574 , 365–371 (2019).

Hodge, R. D. et al. Conserved cell types with divergent features in human versus mouse cortex. Nature 573 , 61–68 (2019).

Hawrylycz, M. J. et al. An anatomically comprehensive atlas of the adult human brain transcriptome. Nature 489 , 391–399 (2012).

Litviňuková, M. et al. Cells of the adult human heart. Nature 588 , 466–472 (2020).

Tucker, N. R. et al. Transcriptional and cellular diversity of the human heart. Circulation 142 , 466–482 (2020).

Giannasca, P. J., Giannasca, K. T., Leichtner, A. M. & Neutra, M. R. Human intestinal M cells display the sialyl Lewis A antigen. Infect. Immun. 67 , 946–953 (1999).

Buettner, M. & Lochner, M. Development and function of secondary and tertiary lymphoid organs in the small intestine and the colon. Front. Immunol. 7 , 342 (2016).

Hoyle, C. H. & Burnstock, G. Neuronal populations in the submucous plexus of the human colon. J. Anat. 166 , 7–22 (1989).

CAS   PubMed   PubMed Central   Google Scholar  

Westerhoff, M. & Greeson, J. in Histology for Pathologists (ed. Mills, S.) Ch. 24, 640–663 (Wolters Kluwer, 2019).

Azzali, G. Structure, lymphatic vascularization and lymphocyte migration in mucosa-associated lymphoid tissue. Immunol. Rev. 195 , 178–189 (2003).

Furness, J. B., Callaghan, B. P., Rivera, L. R. & Cho, H. J. The enteric nervous system and gastrointestinal innervation: integrated local and central control. Adv. Exp. Med. Biol. 817 , 39–71 (2014).

Arai, T. & Kino, I. Morphometrical and cell kinetic studies of normal human colorectal mucosa. Comparison between the proximal and the distal large intestine. Acta Pathol. Jpn 39 , 725–730 (1989).

CAS   PubMed   Google Scholar  

Fenton, T. M. et al. Immune profiling of human gut-associated lymphoid tissue identifies a role for isolated lymphoid follicles in priming of region-specific immunity. Immunity 52 , 557–570 (2020).

Habowski, A. N. et al. Transcriptomic and proteomic signatures of stemness and differentiation in the colon crypt. Commun. Biol. 3 , 453 (2020).

Lundqvist, C., Baranov, V., Hammarström, S., Athlin, L. & Hammarström, M. L. Intra-epithelial lymphocytes. Evidence for regional specialization and extrathymic T cell maturation in the human gut epithelium. Int. Immunol. 7 , 1473–1487 (1995).

Lockyer, M. G. & Petras, R. E. in Histology for Pathologists (ed. Mills, S.) Ch. 25, 664–676 (Wolters Kluwer, 2019).

Pittman, M. E. & Yantiss, R. K. in Histology for Pathologists (ed. Mills, S.) Ch. 26, 677–691 (Wolters Kluwer, 2019).

Kriz, W. & Bankir, L. A standard nomenclature for structures of the kidney. The Renal Commission of the International Union of Physiological Sciences (IUPS). Kidney Int. 33 , 1–7 (1988).

Lake, B. B. et al. A single-nucleus RNA-sequencing pipeline to decipher the molecular anatomy and pathophysiology of human kidneys. Nat. Commun. 10 , 2832 (2019).

Barry, D. M. et al. Molecular determinants of nephron vascular specialization in the kidney. Nat. Commun. 10 , 5705 (2019).

Menon, R. et al. Single cell transcriptomics identifies focal segmental glomerulosclerosis remission endothelial biomarker. JCI Insight 5 , e133267 (2020).

Article   PubMed Central   Google Scholar  

Ransick, A. et al. Single-cell profiling reveals sex, lineage, and regional diversity in the mouse kidney. Dev. Cell 51 , 399–413 (2019).

Limbutara, K., Chou, C. L. & Knepper, M. A. Quantitative proteomics of all 14 renal tubule segments in rat. J. Am. Soc. Nephrol. 31 , 1255–1266 (2020).

Kuppe, C. et al. Decoding myofibroblast origins in human kidney fibrosis. Nature 589 , 281–286 (2021).

Stewart, B. J. et al. Spatiotemporal immune zonation of the human kidney. Science 365 , 1461–1466 (2019).

Kirita, Y., Wu, H., Uchimura, K., Wilson, P. C. & Humphreys, B. D. Cell profiling of mouse acute kidney injury reveals conserved cellular responses to injury. Proc. Natl Acad. Sci. USA 117 , 15874–15883 (2020).

Standring, S. Gray’s Anatomy: The Anatomical Basis of Clinical Practice (Elsevier, 2016).

Haefeli-Bleuer, B. & Weibel, E. R. Morphometry of the human pulmonary acinus. Anat. Rec. 220 , 401–414 (1988).

Whitsett, J. A., Kalin, T. V., Xu, Y. & Kalinichenko, V. V. Building and regenerating the lung cell by cell. Physiol. Rev. 99 , 513–554 (2019).

Plasschaert, L. W. et al. A single-cell atlas of the airway epithelium reveals the CFTR-rich pulmonary ionocyte. Nature 560 , 377–381 (2018).

Xu, Y. et al. Single-cell RNA sequencing identifies diverse roles of epithelial cells in idiopathic pulmonary fibrosis. JCI Insight 1 , e90558 (2016).

Adams, T. S. et al. Single-cell RNA-seq reveals ectopic and aberrant lung-resident cell populations in idiopathic pulmonary fibrosis. Sci. Adv. 6 , eaba1983 (2020).

Wang, A. et al. Single-cell multiomic profiling of human lungs reveals cell-type-specific and age-dynamic control of SARS-CoV2 host genes. eLife 9 , e62522 (2020).

Travaglini, K. J. et al. A molecular cell atlas of the human lung from single-cell RNA sequencing. Nature 587 , 619–625 (2020).

Deprez, M. et al. A single-cell atlas of the human healthy airways. Am. J. Respir. Crit. Care Med. 202 , 1636–1645 (2019).

Medeiros, L. J. et al. in Tumors of the Lymph Node and Spleen Ch. 1 (American Registry of Pathology, 2017).

Medeiros, L. J. et al. Tumors of the Lymph Node and Spleen (American Registry of Pathology, 2017).

O’Malley, D. P., George, T. I., Orazi, A. & Abbondanzo, S. L. in Benign and Reactive Conditions of Lymph Node and Spleen Ch. 1 (American Registry of Pathology, 2009).

Angel, C. E. et al. Distinctive localization of antigen-presenting cells in human lymph nodes. Blood 113 , 1257–1267 (2009).

James, K. R. et al. Distinct microbial and immune niches of the human colon. Nat. Immunol. 21 , 343–353 (2020).

Link, A. et al. Association of T-zone reticular networks and conduits with ectopic lymphoid tissues in mice and humans. Am. J. Pathol. 178 , 1662–1675 (2011).

Park, S. M. et al. Mapping the distinctive populations of lymphatic endothelial cells in different zones of human lymph nodes. PLoS ONE 9 , e106814 (2014).

Xiang, M. et al. A single-cell transcriptional roadmap of the mouse and human lymph node lymphatic vasculature. Front. Cardiovasc. Med. 7 , 52 (2020).

Takeda, A. et al. Single-cell survey of human lymphatics unveils marked endothelial cell heterogeneity and mechanisms of homing for neutrophils. Immunity 51 , 561–572 (2019).

Kunicki, M. A., Hernandez, L. C. A., Davis, K. L., Bacchetta, R. & Roncarolo, M. G. Identity and diversity of human peripheral Th and T regulatory cells defined by single-cell mass cytometry. J. Immunol. 200 , 336–346 (2018).

Pusztaszeri, M. P., Seelentag, W. & Bosman, F. T. Immunohistochemical expression of endothelial markers CD31, CD34, von Willebrand factor, and Fli-1 in normal human tissues. J. Histochem. Cytochem. 54 , 385–395 (2006).

Reynolds, G. et al. Developmental cell programs are co-opted in inflammatory skin disease. Science 371 , eaba6500 (2021).

Dyring-Anderson, B. et al. Spatially and cell-type resolved quantitative proteomic atlas of healthy human skin. Nat. Commun. 11 , 5587 (2020).

Fuchs, E. Keratins and the skin. Annu. Rev. Cell Dev. Biol. 11 , 123–153 (1995).

Nestle, F. O., Meglio, P. D., Qin, J. Z. & Nickoloff, B. J. Skin immune sentinels in health and disease. Nat. Rev. Immunol. 9 , 679–691 (2009).

Laverdet, B. et al. Skin innervation: important roles during normal and pathological cutaneous repair. Histol. Histopathol. 30 , 875–892 (2015).

Ryan, T. J. The blood vessels of the skin. J. Invest. Dermatol. 67 , 110–118 (1976).

Popescu, D. M. A Single Cell Atlas of Adult Healthy, Psoriatic and Atopic Dermatitis Skin (2021); https://developmentcellatlas.ncl.ac.uk/datasets/hca_skin_portal

Bos, J. D. et al. The skin immune system (SIS): distribution and immunophenotype of lymphocyte subpopulations in normal human skin. J. Invest. Dermatol. 88 , 569–573 (1987).

Schweizer, J. et al. New consensus nomenclature for mammalian keratins. J. Cell Biol. 174 , 169–174 (2006).

Eberl, G., Colonna, M., Di Santo, J. P. & McKenzie, A. N. J. Innate lymphoid cells: a new paradigm in immunology. Science 348 , aaa6566 (2015).

Ali, N. & Rosenblum, M. D. Regulatory T cells in skin. Immunology 152 , 372–381 (2017).

Huber, W. E. et al. A tissue-restricted cAMP transcriptional response: SOX10 modulates alpha-melanocyte-stimulating hormone-triggered expression of microphthalmia-associated transcription factor in melanocytes. J. Biol. Chem. 278 , 45224–45230 (2003).

Haniffa, M., Gunawan, M. & Jardine, L. Human skin dendritic cells in health and disease. J. Dermatol. Sci. 77 , 85–92 (2015).

Cesta, M. F. Normal structure, function, and histology of the spleen. Toxicol. Pathol. 34 , 455–465 (2006).

Madissoon, E. et al. scRNA-seq assessment of the human lung, spleen, and esophagus tissue stability after cold preservation. Genome Biol. 21 , 1 (2019).

Pack, M. et al. DEC-205/CD205 + dendritic cells are abundant in the white pulp of the human spleen, including the border region between the red and white pulp. Immunology 123 , 438–446 (2008).

Steiniger, B. S. Human spleen microanatomy: why mice do not suffice. Immunology 145 , 334–346 (2015).

Steiniger, B. S., Seiler, A., Lampp, K., Wilhelmi, V. & Stachniss, V. B lymphocyte compartments in the human splenic red pulp: capillary sheaths and periarteriolar regions. Histochem. Cell Biol. 141 , 507–518 (2014).

Steiniger, B. S., Stachniss, V., Schwarzbach, H. & Barth, P. J. Phenotypic differences between red pulp capillary and sinusoidal endothelia help localizing the open splenic circulation in humans. Histochem. Cell Biol. 128 , 391–398 (2007).

Qiu, J. et al. The characteristics of vessel lining cells in normal spleens and their role in the pathobiology of myelofibrosis. Blood Adv. 2 , 1130–1145 (2018).

Lewis, S. M., Williams, A. & Eisenbarth, S. C. Structure and function of the immune system in the spleen. Sci. Immunol. 4 , eaau6085 (2019).

Mittag, D. et al. Human dendritic cell subsets from spleen and blood are similar in phenotype and function but modified by donor health status. J. Immunol. 186 , 6207–6217 (2011).

Cheng, H. W. et al. Origin and differentiation trajectories of fibroblastic reticular cells in the splenic white pulp. Nat. Commun. 10 , 1739 (2019).

Van Krieken, J. H. J. M. & Te Velde, J. Immunohistology of the human spleen: an inventory of the localization of lymphocyte subpopulations. Histopathology 10 , 285–294 (1986).

Van Krieken, J. H. J. M., Te Velde, J., Leenheers-Binnendijk, L. & Van de Velde, C. J. H. The human spleen; a histological study in splenectomy specimens embedded in methylmethacrylate. Histopathology 9 , 571–585 (1985).

Bautista, J. L. et al. Single-cell transcriptional profiling of human thymic stroma uncovers novel cellular heterogeneity in the thymic medulla. Nat. Commun. 12 , 1096 (2021).

Haynes, B. F. The human thymic microenvironment. Adv. Immunol. 36 , 87–142 (1984).

Park, J. E. et al. A cell atlas of human thymic development defines T cell repertoire formation. Science 367 , eaay3224 (2020).

Pearse, G. Normal structure, function and histology of the thymus. Toxicol. Pathol. 34 , 504–514 (2006).

Suster, S. & Rosai, J. Histology of the normal thymus. Am. J. Surg. Pathol. 14 , 284–303 (1990).

Mignini, F. et al. Neuro-immune modulation of the thymus microenvironment (Review). Int. J. Mol. Med. 33 , 1392–1400 (2014).

Stoeckle, C. et al. Isolation of myeloid dendritic cells and epithelial cells from human thymus. J. Vis. Exp. 79 , e50951 (2013).

Marcovecchio, G. E. et al. Thymic epithelium abnormalities in DiGeorge and Down syndrome patients contribute to dysregulation in T cell development. Front. Immunol. 10 , 447 (2019).

Wakimoto, T. et al. Identification and characterization of human thymic cortical dendritic macrophages that may act as professional scavengers of apoptotic thymocytes. Immunobiology 213 , 837–847 (2008).

Lavaert, M. et al. Integrated scRNA-seq identifies human postnatal thymus seeding progenitors and regulatory dynamics of differentiating immature thymocytes. Immunity 52 , 1088–1104 (2020).

Nuñez, S. et al. The human thymus perivascular space is a functional niche for viral-specific plasma cells. Sci. Immunol. 1 , eaah4447 (2016).

Bendriss-Vermare, N. et al. Human thymus contains IFN-α-producing CD11c − , myeloid CD11c + , and mature interdigitating dendritic cells. J. Clin. Invest. 107 , 835–844 (2001).

Netter, F. H. Atlas of Human Anatomy 7th edn (Elsevier, 2019).

Kandathil, A. & Chamarthy, M. Pulmonary vascular anatomy & anatomical variants. Cardiovasc. Diagn. Ther. 8 , 201–207 (2018).

Perlmutter, D. & Rhoton, A. L. Jr Microsurgical anatomy of the distal anterior cerebral artery. J. Neurosurg. 49 , 204–228 (1978).

Hacein-Bey, L. et al. The ascending pharyngeal artery: branches, anastomoses, and clinical significance. AJNR Am. J. Neuroradiol. 23 , 1246–1256 (2002).

PubMed   PubMed Central   Google Scholar  

Vuong, S. M., Jeong, W. J., Morales, H. & Abruzzo, T. A. Vascular diseases of the spinal cord: infarction, hemorrhage, and venous congestive myelopathy. Semin. Ultrasound CT MR 37 , 466–481 (2016).

Picel, A. C., Hsieh, T. C., Shapiro, R. M., Vezeridis, A. M. & Isaacson, A. J. Prostatic artery embolization for benign prostatic hyperplasia: patient evaluation, anatomy, and technique for successful treatment. Radiographics 39 , 1526–1548 (2019).

Vummidi, D. et al. Pseudolesion in segment IV A of the liver from vein of Sappey secondary to SVC obstruction. Radiol. Case Rep. 5 , 394 (2015).

Tang, F. et al. mRNA-Seq whole-transcriptome analysis of a single cell. Nat. Methods 6 , 377–382 (2009).

Moffitt, J. R. et al. Molecular, spatial, and functional single-cell profiling of the hypothalamic preoptic region. Science https://doi.org/10.1126/science.aau5324 (2018).

International Human Genome Sequencing (IHGS) Consortium Initial sequencing and analysis of the human genome. Nature 409 , 860–921 (2001).

Venter, J. C. et al. The sequence of the human genome. Science 291 , 1304–1351 (2001).

Visible Human Project (VHP) Data Sets (NLM, 2020); https://www.nlm.nih.gov/databases/download/vhp.html

Fair Principles (Go Fair, 2021); https://www.go-fair.org/fair-principles

CCF Anatomical Structures, Cell Types and Biomarkers (ASCT+B) Tables (HuBMAP Consortium, 2021); https://hubmapconsortium.github.io/ccf/pages/ccf-anatomical-structures.html

CCF 3D Reference Object Library (HuBMAP Consortium, 2021); https://hubmapconsortium.github.io/ccf/pages/ccf-3d-reference-library.html

Graphics Language Transmission Format (glTF) a File Format Specification for 3D Scenes and Model s (Khronos Group, 2021); https://www.khronos.org

CCF Registration User Interface (HuBMAP Consortium, 2021); https://hubmapconsortium.github.io/ccf-ui/rui

CCF Exploration User Interface (HuBMAP Consortium, 2021); https://portal.hubmapconsortium.org/ccf-eui

Human Biomolecular Atlas Program (HuBMAP) Data Portal (NIH, 2020); https://portal.hubmapconsortium.org

CCF User Interfaces (RUI, EUI) (HuBMAP Consortium, 2021); https://github.com/hubmapconsortium/ccf-ui

Jardine, L. et al. Blood and immune development in human fetal bone marrow and Down syndrome. Nature 598 , 327–331 (2021).

Stephenson, E. et al. Single-cell multi-omics analysis of the immune response in COVID-19. Nat. Med. 27 , 904–916 (2021).

Lake, B. B. et al. An atlas of healthy and injured cell states and niches in the human kidney. Preprint at bioRxiv https://doi.org/10.1101/2021.07.28.454201 (2021).

Zhuo, L., Huang, L., Yang, Z., Li, G. & Wang, L. A comprehensive analysis of NPHS1 gene mutations in patients with sporadic focal segmental glomerulosclerosis. BMC Med. Genet. 20 , 111 (2019).

Ong, E. et al. Modelling kidney disease using ontology: insights from the Kidney Precision Medicine Project. Nat. Rev. Nephrol. 16 , 686–696 (2020).

Barwinska, D. et al. Molecular characterization of the human kidney interstitium in health and disease. Sci. Adv. 7 , eabd3359 (2021).

ASCT+B Reporter (HuBMAP Consortium, 2021); https://hubmapconsortium.github.io/ccf-asct-reporter

Majumder, P. P., Mhlanga, M. M. & Shalek, A. K. The Human Cell Atlas and equity: lessons learned. Nat. Med. 26 , 1509–1511 (2020).

Spatial Biology Europe: Online (HuBMAP Consortium, 2021); https://cns-iu.github.io/workshops/2021-2004-2014_spatial_biology_europe_online

Azimuth App for reference-based single-cell analysis (Satija Lab, 2021); https://azimuth.hubmapconsortium.org/

Hao, Y. et al. Integrated analysis of multimodal single-cell data. Cell 184 , 3573–3587 (2021).

Osumi-Sutherland, D. et al. Cell type ontologies of the Human Cell Atlas. Nat. Cell Biol. https://doi.org/10.1038/s41556-021-00787-7 (2021).

Radtke, A. J. et al. IBEX: an open and extensible method for high content multiplex imaging of diverse tissues. Preprint at https://arxiv.org/abs/2107.11364 (2021).

Hickey, J. W. et al. Spatial mapping of protein composition and tissue organization: a primer for multiplexed antibody-based imaging. Preprint at https://arxiv.org/abs/2107.07953 (2021).

Manz, T. et al. Viv: multiscale visualization of high-resolution multiplexed bioimaging data on the web. Preprint at OSF https://doi.org/10.31219/osf.io/wd2gu (2020).

Balhoff, J. & Curtis, C. K. Ubergraph (2021); https://github.com/INCATools/ubergraph

Li, W., Germain, R. N. & Gerner, M. Y. Multiplex, quantitative cellular analysis in large tissue volumes with clearing-enhanced 3D microscopy (Ce3D). Proc. Natl Acad. Sci. USA 114 , E7321–E7330 (2017).

Fedorov, A. et al. 3D Slicer as an image computing platform for the Quantitative Imaging Network. Magn. Reson. Imaging 30 , 1323–1341 (2012).

Slicer Community. 3D Slicer image computing platform (2021); https://www.slicer.org/

ZBrush, the all-in-one digital sculpting solution (Pixologic, 2021); https://pixologic.com/

Maya: 3D computer animation, modeling, simulation, and rendering software (Autodesk, 2021); https://www.autodesk.com/products/maya/overview

Download references

Acknowledgements

We thank B. B. Lake from University of California, San Diego, for assistance with annotations and analysing the single-nucleus RNA-seq HUBMAP data for several of the markers in the kidney ASCT+B tables; B. Steck and R. Dull from the University of Michigan for assistance with the nomenclature and curation of kidney partonomy; S. Winfree, IUPUI, for discussions regarding the kidney ASCT+B table; and staff at the KPMP, especially the Tissue Interrogation Sites and the Controlled Cell Vocabulary working group, for guidance and development of the initial sets of ASCT+B kidney tables. We acknowledge L. Yao for segmenting and optimizing the mouse popliteal lymph node model from high-resolution microscopy data. The work was funded, in part, by NIH Awards OT2OD026671, U54DK120058, 1UH3CA246594, 1U54AI142766, 1UG3CA256960, 1UG3HL145609, U54HL145608, U54HL145611, UH3DK114933, DK110814 and DK107350; National Institute of Allergy and Infectious Diseases (NIAID), Department of Health and Human Services under BCBB Support Services Contract HHSN316201300006W/HHSN27200002; the Intramural Research Program of the NIH at NIAID; and Helmsley Charitable Trust 2018PG-T1D071.

Author information

Authors and affiliations.

Department of Intelligent Systems Engineering, Indiana University, Bloomington, IN, USA

Katy Börner, Ellen M. Quardokus, Bruce W. Herr II, Andreas Bueckle & Hrishikesh Paul

Wellcome Sanger Institute, Wellcome Genome Campus, Hinxton, Cambridge, UK

Sarah A. Teichmann

Department of Radiology, University of Pennsylvania, Philadelphia, PA, USA

James C. Gee

Department of Health and Human Services, National Institute of Allergy and Infectious Diseases, National Institutes of Health, Bethesda, MD, USA

Kristen Browne

European Bioinformatics Institute (EMBL-EBI), Wellcome Trust Genome Campus, Cambridge, UK

David Osumi-Sutherland

Biosciences Institute, Newcastle University, Newcastle upon Tyne, UK

Muzlifah Haniffa & Laura Jardine

The Kavli Foundation, Los Angeles, CA, USA

Amy Bernard

Allen Institute, Seattle, WA, USA

Song-Lin Ding & Jeremy A. Miller

Department of Medicine, University of Washington, Seattle, WA, USA

Department of Pathology, Johns Hopkins University School of Medicine, Baltimore, MD, USA

Marc K. Halushka

Department of Computer Science, Princeton University, Princeton, NJ, USA

Avinash Boppana

Department of Microbiology and Immunology, Stanford University School of Medicine, Stanford, CA, USA

Teri A. Longacre & John Hickey

Department of Surgery, Washington University in St Louis, St Louis, MO, USA

Harvard Institute of Medicine, Harvard Medical School, Boston, MA, USA

M. Todd Valerius & Rajeev Malhotra

Department of Microbiology and Immunology, and Center for Computational Medicine and Bioinformatics, University of Michigan Medical School, Ann Arbor, MI, USA

Department of Pediatrics, University of Rochester, Rochester, NY, USA

Gloria Pryhuber

Biological Sciences, University of California, San Diego, La Jolla, CA, USA

Department of Pathology, Immunology and Laboratory Medicine, University of Florida, Gainesville, FL, USA

Marda Jorgensen, Clive Wasserfall & Maigan Brusko

Center for Advanced Tissue Imaging, Laboratory of Immune System Biology, NIAID, NIH, Bethesda, MD, USA

Andrea J. Radtke & Rebecca T. Beuschel

Biology and Applied Physics, General Electric Research, Niskayuna, NY, USA

Fiona Ginty

Department of Dermatology, University of Pittsburgh, Pittsburgh, PA, USA

Department of Dermatology, Johns Hopkins School of Medicine, Baltimore, MD, USA

Joel Sunshine

Division of Vascular Surgery and Endovascular Therapy, Massachusetts General Hospital, Boston, MA, USA

Sujin Lee & Rajeev Malhotra

Department of Medicine, Washington University School of Medicine, St Louis, MO, USA

Sanjay Jain

Department of Pathology and Immunology, Washington University School of Medicine, St Louis, MO, USA

Department of Biomedical Informatics, Harvard Medical School, Boston, MA, USA

Griffin Weber

You can also search for this author in PubMed   Google Scholar

Corresponding author

Correspondence to Katy Börner .

Ethics declarations

Competing interests.

In the past 3 years, S.A.T. has received remuneration for consulting and Scientific Advisory Board membership from Genentech, Roche, Biogen, GlaxoSmithKline, Foresite labs, Qiagen and Transition Bio and she is a co-founder and equity holder of Transition Bio. R.M. receives research funding from Bayer and Amgen and serves as a consultant for Myokardia/BMS and Third Pole; he is a co-founder of Patch Inc; he is a co-inventor for a patent no. PCT/US2O12/O22119 on pharmacologic BMP inhibitors (along with Mass General Brigham) for which he receives royalties from Keros Therapeutics, Inc.; he also receives royalties from UpToDate for scientific content authorship. The other authors declare no competing interests.

Additional information

Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Supplementary information

Supplementary information.

ASCT+B Tables

Rights and permissions

Reprints and permissions

About this article

Cite this article.

Börner, K., Teichmann, S.A., Quardokus, E.M. et al. Anatomical structures, cell types and biomarkers of the Human Reference Atlas. Nat Cell Biol 23 , 1117–1128 (2021). https://doi.org/10.1038/s41556-021-00788-6

Download citation

Received : 30 January 2021

Accepted : 29 September 2021

Published : 08 November 2021

Issue Date : November 2021

DOI : https://doi.org/10.1038/s41556-021-00788-6

Share this article

Anyone you share the following link with will be able to read this content:

Sorry, a shareable link is not currently available for this article.

Provided by the Springer Nature SharedIt content-sharing initiative

This article is cited by

Transformers in single-cell omics: a review and new perspectives.

  • Artur Szałata
  • Karin Hrovatin
  • Fabian J. Theis

Nature Methods (2024)

Publication, funding, and experimental data in support of Human Reference Atlas construction and usage

  • Yongxin Kong
  • Katy Börner

Scientific Data (2024)

Dictionary learning for integrative, multimodal and scalable single-cell analysis

  • Rahul Satija

Nature Biotechnology (2023)

Segmentation of human functional tissue units in support of a Human Reference Atlas

  • Yashvardhan Jain
  • Leah L. Godwin

Communications Biology (2023)

AtOM, an ontology model to standardize use of brain atlases in tools, workflows, and data infrastructures

  • Heidi Kleven
  • Thomas H. Gillespie
  • Trygve B. Leergaard

Scientific Data (2023)

Quick links

  • Explore articles by subject
  • Guide to authors
  • Editorial policies

Sign up for the Nature Briefing newsletter — what matters in science, free to your inbox daily.

research paper on brain anatomy

U.S. flag

An official website of the United States government

The .gov means it’s official. Federal government websites often end in .gov or .mil. Before sharing sensitive information, make sure you’re on a federal government site.

The site is secure. The https:// ensures that you are connecting to the official website and that any information you provide is encrypted and transmitted securely.

  • Publications
  • Account settings
  • My Bibliography
  • Collections
  • Citation manager

Save citation to file

Email citation, add to collections.

  • Create a new collection
  • Add to an existing collection

Add to My Bibliography

Your saved search, create a file for external citation management software, your rss feed, anatomy, central nervous system, affiliations.

  • 1 McKinsey & Company
  • 2 Augusta University Medical Center
  • PMID: 31194336
  • Bookshelf ID: NBK542179

The nervous system is divided into the central nervous system (CNS) and the peripheral nervous system. The CNS includes the brain and spinal cord, while the peripheral nervous system consists of everything else. The CNS's responsibilities include receiving, processing, and responding to sensory information (see Image. Peripheral and Central Nervous Systems).

The brain is an organ of nervous tissue responsible for responses, sensation, movement, emotions, communication, thought processing, and memory. The skull, meninges, and cerebrospinal fluids protect the human brain. The nervous tissue is extremely delicate and can be damaged by the smallest amount of force. In addition, the brain has a blood-brain barrier that prevents the brain from any harmful substance floating in the blood.

The spinal cord is a vital aspect of the CNS found within the vertebral column. Its purpose is to send motor commands from the brain to the peripheral body and relay sensory information from the sensory organs to the brain. Bone, meninges, and cerebrospinal fluids provide spinal cord protection.

Copyright © 2024, StatPearls Publishing LLC.

PubMed Disclaimer

Conflict of interest statement

Disclosure: Lauren Thau declares no relevant financial relationships with ineligible companies.

Disclosure: Vamsi Reddy declares no relevant financial relationships with ineligible companies.

Disclosure: Paramvir Singh declares no relevant financial relationships with ineligible companies.

  • Introduction
  • Structure and Function
  • Blood Supply and Lymphatics
  • Surgical Considerations
  • Clinical Significance
  • Review Questions

Similar articles

  • Anatomy, Back, Spinal Cord Arteries. Kaiser JT, Reddy V, Lugo-Pico JG. Kaiser JT, et al. 2023 Aug 14. In: StatPearls [Internet]. Treasure Island (FL): StatPearls Publishing; 2024 Jan–. 2023 Aug 14. In: StatPearls [Internet]. Treasure Island (FL): StatPearls Publishing; 2024 Jan–. PMID: 30725904 Free Books & Documents.
  • Applications of Proteomics to Nerve Regeneration Research. Massing MW, Robinson GA, Marx CE, Alzate O, Madison RD. Massing MW, et al. In: Alzate O, editor. Neuroproteomics. Boca Raton (FL): CRC Press/Taylor & Francis; 2010. Chapter 15. In: Alzate O, editor. Neuroproteomics. Boca Raton (FL): CRC Press/Taylor & Francis; 2010. Chapter 15. PMID: 21882439 Free Books & Documents. Review.
  • Neuroanatomy, Somatic Nervous System. Akinrodoye MA, Lui F. Akinrodoye MA, et al. 2022 Nov 7. In: StatPearls [Internet]. Treasure Island (FL): StatPearls Publishing; 2024 Jan–. 2022 Nov 7. In: StatPearls [Internet]. Treasure Island (FL): StatPearls Publishing; 2024 Jan–. PMID: 32310487 Free Books & Documents.
  • Neuroanatomy, Spinal Cord. Khan YS, Lui F. Khan YS, et al. 2023 Jul 24. In: StatPearls [Internet]. Treasure Island (FL): StatPearls Publishing; 2024 Jan–. 2023 Jul 24. In: StatPearls [Internet]. Treasure Island (FL): StatPearls Publishing; 2024 Jan–. PMID: 32644482 Free Books & Documents.
  • Barrier function in the peripheral and central nervous system-a review. Reinhold AK, Rittner HL. Reinhold AK, et al. Pflugers Arch. 2017 Jan;469(1):123-134. doi: 10.1007/s00424-016-1920-8. Epub 2016 Dec 12. Pflugers Arch. 2017. PMID: 27957611 Review.
  • Shipp S. Structure and function of the cerebral cortex. Curr Biol. 2007 Jun 19;17(12):R443-9. - PubMed
  • Lanciego JL, Luquin N, Obeso JA. Functional neuroanatomy of the basal ganglia. Cold Spring Harb Perspect Med. 2012 Dec 01;2(12):a009621. - PMC - PubMed
  • Flament-Durand J. The hypothalamus: anatomy and functions. Acta Psychiatr Belg. 1980 Jul-Aug;80(4):364-75. - PubMed
  • Nattie E, Li A. Central chemoreceptors: locations and functions. Compr Physiol. 2012 Jan;2(1):221-54. - PMC - PubMed
  • Rajmohan V, Mohandas E. The limbic system. Indian J Psychiatry. 2007 Apr;49(2):132-9. - PMC - PubMed

Publication types

  • Search in PubMed
  • Search in MeSH
  • Add to Search

Related information

  • Cited in Books

LinkOut - more resources

Full text sources.

  • NCBI Bookshelf

book cover photo

  • Citation Manager

NCBI Literature Resources

MeSH PMC Bookshelf Disclaimer

The PubMed wordmark and PubMed logo are registered trademarks of the U.S. Department of Health and Human Services (HHS). Unauthorized use of these marks is strictly prohibited.

U.S. flag

An official website of the United States government

The .gov means it’s official. Federal government websites often end in .gov or .mil. Before sharing sensitive information, make sure you’re on a federal government site.

The site is secure. The https:// ensures that you are connecting to the official website and that any information you provide is encrypted and transmitted securely.

  • Publications
  • Account settings

Preview improvements coming to the PMC website in October 2024. Learn More or Try it out now .

  • Advanced Search
  • Journal List
  • v.7(9); 2006 Sep

Logo of emborep

Inside the brain of a neuron

Kyriaki sidiropoulou.

1 Institute of Molecular Biology and Biotechnology (IMBB), Foundation for Research and Technology–Hellas (FORTH), Vassilika Vouton PO Box 1583, Heraklion GR71110, Crete, Greece

Eleftheria Kyriaki Pissadaki

2 Department of Biology, University of Crete, Vassilika Vouton, Heraklion GR71409, Crete, Greece

Panayiota Poirazi

For many decades, neurons were considered to be the elementary computational units of the brain and were assumed to summate incoming signals and elicit action potentials only in response to suprathreshold stimuli. Although modelling studies predicted that single neurons constitute a much more powerful computational entity, able to perform an array of nonlinear calculations, this possibility was not explored experimentally until the discovery of active mechanisms in the dendrites of most neuron types. Here, we review several modelling studies that have addressed information processing in single neurons, starting with those characterizing the arithmetic of different dendritic components, to those tackling neuronal integration at the cell body and, finally, those analysing the computational abilities of the axon. We present modelling predictions along with supporting experimental data in an effort to highlight the significant contribution of modelling work to enhancing our understanding of single-neuron arithmetic.

Introduction

Understanding how the brain works remains one of the most exciting and intricate challenges of modern biology. Despite the wealth of information that has accumulated during the past years about the molecular and biophysical mechanisms that underlie neuronal activity, similar advances have yet to be made in understanding the rules that govern information processing and the relationship between the structure and function of a neuron.

Computational models provide a theoretical framework together with a technological platform for enhancing our understanding of nervous system functions. Certain tools are suitable for efficiently analysing and interpreting complex data sets, such as multi-channel recordings from hundreds of neurons, whereas others are used to simulate the activity of single cells, neural networks or systems of networks at various levels of abstraction. The development and application of such modelling tools enable researchers to quantitatively investigate several hypotheses by using interactive models of the systems under study. When used in conjunction with experimental techniques, these models facilitate hypothesis testing and help to identify key follow-up experiments.

In this review, we discuss several computational studies in which realistic biophysical models have been used to elucidate the computational tasks performed by a neuron. We focus on single-neuron models that incorporate a significant level of detail and compare modelling predictions with experimental findings. Although a great amount of work has also been devoted to modelling neural components as well as neuronal assemblies at a more abstract level, reviewing these studies is not the purpose of this article.

Single-neuron computations

Whether incredibly simple as bipolar cells in the retina or immensely complex as Purkinje cells in the cerebellum ( Ramon y Cajal, 1933 ), most neurons are composed of three main structural units: the dendrites, the soma (cell body) and the axon. For the past few decades, axons and dendrites have been considered to be simple transmitting devices that communicate signals to and from the soma in which thresholded computations take place. As a result, neuronal cells were initially represented as spherical point neurons—consisting only of a cell body—and information transfer was thought to lie entirely in their average firing rates ( McCulloch & Pitts, 1943 ). However, primarily computational, and more recently physiological, studies have shown that variations in the morphology and ionic conductance composition of different neurons provide the cell with enhanced computational abilities far outreaching those captured by a point neuron.

Computing with dendrites: new roles for old structures

The old view that dendrites are merely passive cables that relay incoming signals to the cell body no longer holds true. In the light of accumulating evidence highlighting the active role of dendrites in signal integration, these structures seem able to perform a variety of computational tasks, including temporal integration, signal amplification and attenuation, and detection of coincident incoming inputs (for a recent review, see London & Hausser, 2005 ). In this section, we further elucidate the role of dendrites in the information processing capacity of the neuron by focusing on insights gained primarily from modelling studies and by using a bottom-up approach: starting from the smallest dendritic subunit—the spine—up to the effect of network activity on dendritic and, subsequently, neuronal function.

Computations carried out by excitable spines

Dendritic spines were anatomically identified by Ramon y Cajal in 1911, who referred to them as espinas due to their resemblance to thorns on flower stems (for a review, see Segal, 2002 ). Theoretical findings first indicated that the anatomical characteristics of spines, as well as the possible presence of voltage-gated ion channels, allow for compartmentalized gain modulation of synaptic inputs in spine heads—that is, information can be combined nonlinearly from two or more sources ( Segev & Rall, 1998 ). Models predicted that strong inputs are able to initiate local dendritic spikes ( Baer & Rinzel, 1991 ), which in turn could activate additional nearby spines ( Shepherd et al , 1985 ) and therefore locally amplify incoming inputs. As schematically depicted in Fig 1 , individual spines can perform nonlinear integration of incoming coincident signals, and their interaction results in a spatially restricted enhancement of dendritic events. By contrast, if synaptic stimulation is sparse, the added resistance and capacitance load provided by the spine membrane coupled with the narrow spine neck would act as a local filter, promoting linear integration ( Yuste & Urban, 2004 ). In a recent compartmental modelling study, spine geometry together with a high density of sodium ion (Na + ) channels on spines might explain the efficacy of somatic action potentials for invading apical dendrites in CA1 pyramids ( Tsay & Yuste, 2002 , 2004 ). This implies a role for these structures in controlling back-propagating signals and perhaps in coincidence detection ( Tsay & Yuste, 2002 , 2004 ). Taken together, these studies reveal the exciting possibility that dendritic sections containing small clusters of spines act as individual computational units. Interestingly, in a recent experimental study, short (around 40 μm) basal dendritic compartments in layer V pyramidal neurons were shown to summate local signals as individually thresholded sigmoidal units ( Polsky et al , 2004 ), which supports this hypothesis.

An external file that holds a picture, illustration, etc.
Object name is 7400789-f1.jpg

The multiplicative neuron. Sigmoidal functions indicate nonlinear computations performed by various parts of the cell: gain modulation in spine heads, thresholded computations in small dendritic branchlets, and supralinearities at the main apical trunk and the cell body. The response of the cell is tuned by the overall effect of the network.

Dendritic computations

If sparse and clustered synaptic inputs can be differentially sensed by a dendritic section, is it possible that different spatial arrangements of synaptic inputs are differentially perceived by the cell? Early neuron models incorporating passive dendritic properties and applying the cable theory—that is, how voltage changes are propagated along dendritic segments—indicated a possible linear summation of inputs arriving at separate parts of the dendritic tree ( Rall, 1959 ). This implied that location is not important. By contrast, inputs that are close together were thought to combine sublinearly due to the activation of a shunting current ( Rall et al , 1967 ). Despite the simplicity of the early models and the presence of various active membrane mechanisms that could in principle support supralinear dendritic integration, experimental evidence in various neuron types has mostly reinforced the linear or sublinear summation rule ( Cash & Yuste, 1999 ; Magee & Cook, 2000 ; Tamas et al , 2002 ). However, at least two studies have shown the presence of powerful thresholding events isolated in the thin dendrites of neocortical ( Schiller et al , 2000 ) and CA1 ( Wei et al , 2001 ) pyramidal cells.

This idea of spatial compartmentalization in the neuron and its role in information processing was explored further with the use of a detailed CA1 pyramidal neuron model ( Poirazi et al , 2003a , b ). According to the model, each apical oblique dendrite—or part of it—acts as an independent computational unit that summates inputs using a sigmoidal activation function. Different branch outputs are then linearly combined at the cell body. Both of these predictions were verified experimentally in a layer V pyramidal neuron ( Polsky et al , 2004 ), and a recent study confirmed that radial oblique dendrites of CA1 pyramidal neurons function as single integrative compartments ( Losonczy & Magee, 2006 ). As shown in Fig 2 , layer V neocortical neurons linearly summate between-branch excitatory postsynaptic potentials (EPSPs), but implement a sigmoidal activation function for within-branch EPSPs ( Fig 2B ), as predicted by models of CA1 neurons ( Fig 2A ). In other words, thin dendritic branches seem to be able to combine incoming signals according to a thresholding nonlinearity, in a similar way to a typical point neuron. Interestingly, when synaptic inputs vary in both their temporal and spatial distribution, the distal apical trunk of a CA1 pyramidal cell operates in two fundamentally distinct integration forms ( Gasparini & Magee, 2006 ). Asynchronous or spatially distributed synaptic inputs—similar to those occurring during theta oscillations—summate linearly. Synchronous and clustered inputs—similar to those occurring during sharp waves—summate according to a steep sigmoidal nonlinearity. This bimodal dendritic integration code allows a single cell to perform two different state-dependent computations: input strength encoding during theta states and feature detection during sharp waves ( Gasparini & Magee, 2006 ).

An external file that holds a picture, illustration, etc.
Object name is 7400789-f2.jpg

Nonlinear dendritic computations. ( A ) Summation of paired, single-pulse inputs in the apical dendrites of a CA1 pyramidal model cell. Simulations predict a sigmoidal modulation of combined excitatory postsynaptic potentials (EPSPs) within a branch (red symbols) and a linear summation of EPSPs between branches (green symbols). Red curves correspond to within-branch data for dendrites at 92 μm (short dashes), 232 μm (solid) and 301 μm (long dashes) from the soma. Due to differences in local compared with somatic responses, axis values for the red curves are scaled up by a factor of 10. ( B ) Experimental verification of predicted summation rules in basal dendrites of a layer V pyramidal neuron. Single-pulse stimulation of synapses in different branches results in linear summation at the cell body (green squares). When γ-aminobutyric acid (GABA)ergic inhibition is blocked, the summation of within-branch EPSPs is modulated by a sigmoidal nonlinearity (all other symbols). Reproduced with permission from Polsky et al (2004) . ( C ) Schematic representation of a pyramidal neuron as a two-layer neural network. Radial oblique dendrites provide the first layer of the network, each performing individually thresholded computations as shown in ( A ) and ( B ). The outputs of this layer feed into the cell body, which constitutes the second layer of the network model. Adapted from Poirazi et al (2003b) .

Computing with regenerative dendritic events

What would be the benefit of a cell consisting of dendrites that are able to isolate complex events? Early physiological studies identified the presence of active dendritic events in neurons ( Kandel & Spencer, 1961 ); however, they were not further investigated experimentally at that time. Modelling studies predicted a role for Na + -mediated dendritic spikes—initially modelled with H-H conductances—in allowing the back-propagation of action potentials into the dendritic tree ( Rall & Shepherd, 1968 ), as well as enabling coincidence detection between proximal and distal inputs ( Softky & Koch, 1993 ).

A closer evaluation of dendritic events revealed that these spikes can be mediated by Na + ( Golding & Spruston, 1998 ) or calcium ion (Ca 2+ ; Yuste et al , 1994 ) channels in the distal dendrites, as well as by N-methyl D-aspartate (NMDA) channels ( Schiller et al , 2000 ) or a combination of Na + and NMDA channels ( Ariav et al , 2003 ) in the basal dendrites of pyramidal neurons. The quest is now to reveal their role in neuronal function. Although these spikes can be confined to their dendritic site of origin ( Wei et al , 2001 ; Schiller et al , 2000 ; Zhu, 2000 ), physiologically relevant situations such as large, suprathreshold synaptic stimuli in the distal dendrites or the activation of several branches together allow these spikes to act globally and modulate neuronal output ( Zhu, 2000 ; Larkum et al , 2001 ). In a compartmental CA1 neuron model, dendritic spike initiation in response to perforant path (PP) stimulation, which is confined in the distal tuft, could be transferred to the soma by activation of the Schaffer collateral (SC) pathway ( Jarsky et al , 2005 ). Similarly, dendritic Ca 2+ conductances in the distal tuft of a layer V cortical neuron model, which are unable to initiate an action potential at the soma, could be amplified by coincident back-propagating action potentials ( Larkum et al , 2004 ); in addition, dendritic Ca 2+ spikes have been suggested to convert single spikes into bursts in CA3 pyramidal neuron models ( Traub et al , 1991 ), leading to an enhanced neuronal response.

Although the above studies indicate that dendritic events might amplify neuronal gain and facilitate coincident detection of inputs, simulated experiments in a detailed CA1 pyramidal neuron model indicate that distal dendritic activation could bidirectionally gate suprathreshold SC input (E.K.P. and P.P., unpublished data). In particular, PP theta-burst stimulation coincident with or slightly preceding regular SC input, initiates Ca 2+ spikes in the distal dendrites and transforms a previously regular firing response into bursting ( Fig 3A,B ). By contrast, when subthreshold PP stimulation precedes the SC input by a few hundreds of milliseconds, the distal dendrites are hyperpolarized due to enhanced γ-aminobutyric acid (GABA) transmission. This results in a reduced gain of neuronal output, as seen by the blocking of the spikes induced by SC input only ( Fig 3A,C ). This PP-induced ‘spike blocking' was previously reported experimentally ( Dvorak-Carbone & Schuman, 1999 ). Collectively, dendritic regenerative events provide a means by which several localized events are combined to modulate neuronal output, thus expanding the response flexibility of single neurons.

An external file that holds a picture, illustration, etc.
Object name is 7400789-f3.jpg

Top-down gain modulation in a CA1 pyramidal neuron model. In a CA1 pyramidal neuron, afferents from the entorhinal cortex (EC) synapse onto distal dendrites, whereas axons from CA3 neurons project to proximal dendrites. ( A ) Synaptic activation (1 Hz) of proximal dendrites induces a regular firing response at the cell body of the model cell. ( B ) Strong (theta-burst) stimulation of EC inputs coinciding with CA3 stimulation evokes calcium spikes in the distal dendrites, the occurrence of which initiates bursts of action potentials (APs), and facilitates the somatic response. ( C ) Temporally distant activation of EC with subthreshold bursts (400 ms after the CA3 input) prevents the initiation of somatic action potentials for several seconds, thus weakening the somatic response. Green dots, excitatory inputs; red dots, inhibitory inputs.

Normalizing effects on synaptic integration

Whereas dendritic regenerative events allow gain modulation of neuronal output, passive properties and K + conductances in the dendrites act to spatially normalize and temporally integrate inputs, which provides the neuron with a different mode of information processing.

These passive properties (such as membrane time constant, input resistance and dendritic length) allow for differential integration of synaptic inputs that arrive at distal or proximal parts of the dendritic tree owing to the ‘large voltage attenuation and significant temporal delay' of propagated signals ( Koch & Segev, 2000 ). At the same time, incorporation of a hyperpolarization-activated, potassium ion (K + ) conductance ( I h ) in a CA1 compartmental model was shown to account for the experimentally observed normalization of EPSPs that originate from different parts of the dendritic tree so that all inputs induce similar depolarizations at the cell body ( Golding et al , 2005 ; Magee & Cook, 2000 ). Additional modelling experiments indicated that I h might be involved in setting the temporal window for input summation around subthreshold levels, thus enabling coincidence detection and minimizing the effectiveness of non-synchronized inputs ( Migliore et al , 2004 ; Migliore, 2003 ). Other dendritic K + conductances could either support or counteract the effect of I h on temporal summation ( Day et al , 2005 ).

Finally, the long-standing but quite neglected effect of background noise due to network activity on neuronal information processing should be considered. According to the work of Bernander and colleagues (1991) , background activity in a passive neuron model dampens the effectiveness of asynchronous—but not synchronous—inputs in generating a somatic action potential. This in turn facilitates the distinction between ‘unimportant' and ‘meaningful' signals, respectively. In a more detailed cortical neuron model that incorporates active dendritic conductances, intense synaptic network activity was shown to increase the membrane conductance, promote the location-independent effect of inputs arriving onto different dendritic regions (but see London & Segev, 2001 ) and increase the probability for dendritic spike initiation and its forward propagation to the axon ( Rudolph & Destexhe, 2003 ). Thus, incoming background noise from network activity could greatly influence the integrative properties of a neuron—for example, by modulating the spatiotemporal window for dendritic nonlinearities ( Azouz, 2005 ).

In conclusion, the neuron does not behave as a point neuron, but it might consist of many different point neurons in the form of a cluster of spines or a stretch of dendrite, each of which has its own integration rules according to its spatial location and temporal architecture of incoming inputs. When locally induced signals manage to escape their subunit, they are affected by global cellular parameters and are set by the overall network activity the cell receives, promoting a quasi-linear interaction mode.

The overall picture emerging from this analysis is that a single neuron could be decomposed into a multi-layer neural network, able to perform all sorts of nonlinear computations ( London & Hausser, 2005 ). Interestingly, the average firing rate of a detailed CA1 model to hundreds of different input patterns was accurately predicted by a two-layer neural network abstraction ( Fig 2 ), in which individual oblique dendrites provided the first layer and the soma acted as the output layer ( Poirazi et al , 2003b ). This implies a much larger storage capacity than originally assumed for single neurons. According to another computational study, the pattern discrimination capacity of such a cell exceeds that of a point neuron by at least one order of magnitude ( Poirazi & Mel, 2001 ). An even more complex single-neuron unit proposed by Hausser & Mel (2003) entails a two-compartment model in which the distal tuft acts as one compartment and the thin branches of the perisoma act as the other. Both compartments act as two-layer networks whose outputs combine at the cell body, giving rise to an extra powerful, three-layer computing unit.

Information processing at the cell body

Dendrites contribute to nonlinear summation of inputs, whereas the soma might support a different kind of information processing—that of enabling a persistent firing mode in the absence of stimulation. Recent experimental and modelling studies have highlighted the importance of somatic intrinsic membrane mechanisms in generating and maintaining persistent activity, in addition to the traditional network mechanisms (for a review, see Major & Tank, 2004 ). In vitro work in the entorhinal cortex ( Egorov et al, 2002 ; Tahvildari & Alonso, 2005 ) showed that a single neuron is able to generate graded persistent activity under pharmacological stimulation of muscarinic acetylcholine receptors in response to somatic or synaptic stimulation, owing to activation of a slow Ca 2+ -dependent mixed ionic (CAN) conductance. Modelling work has emphasized a possible involvement of the slow temporal decay of the EPSP ( Lisman et al , 1998 ), the CAN conductance ( Tegner et al , 2002 ) or the Ca 2+ -induced Ca 2+ release mechanism ( Loewenstein & Sompolinsky, 2003 ) in the maintenance of a stable persistent state at low physiological frequencies (10–50 Hz).

Persistent activity in vivo has also been observed in the hippocampus ( Wirth et al , 2003 ), although the underlying mechanisms are unclear. Ongoing work in our laboratory shows that persistent activity in a detailed CA1 pyramidal neuron model can be induced in response to theta-burst synaptic stimulation as well as in response to somatic stimulation under the influence of cholinergic modulation. The model supports a role for a slow, Ca 2+ -dependent tail current in maintaining sustained activity in hippocampal neurons ( Poirazi, 2005 ). The ability of single cells to be persistently active further enhances their computational power by adding another mode to their repertoire of complex functions.

Computing with axons

Axons provide a medium through which information, in the form of action potentials, flows across neuronal assemblies. Several computational studies have provided insights into the ionic mechanism for action-potential generation, particularly the seminal work of Hodgkin and Huxley, as well as the action-potential initiation site (reviewed in Stuart et al , 1997 ). More recent studies focusing on the reliability and accuracy of action-potential generation and propagation indicate that these properties, which are crucial for normal information processing, could be modified by changes in axonal geometry and ionic conductance composition ( Segev & Schneidman, 1999 ; Debanne, 2004 ). The work of Goldstein & Rall (1974) in simulated axons with changing diameter and different branching patterns first implicated the above structural characteristics in modifying action-potential curve and propagation speed. After quantifying the effect of such morphological changes on action-potential propagation velocity, another computational study suggested that synaptic boutons in different terminals are activated asynchronously in a reconstructed axon of a layer V neuron in the somatosensory cortex ( Manor et al , 1991 ). The density and type of ionic channels along the axon and branching points has also been suggested to gate action-potential propagation. For example, activation of just a few clusters of A-type K + channels has been shown to gate axonal propagation of action potentials in CA3 neurons in both simulations and experiments ( Debanne et al , 1997 ; Kopysova & Debanne, 1998 ). When combined, these findings imply that axons are much more than simple conducting devices for action potentials. Enriched with a variety of computational abilities, axons seem to be complex transmitting devices whose role in neuronal information processing is worth investigating thoroughly.

Concluding remarks

For several years, neuroscientists believed that the brain's transistor or fundamental processing unit was the neuron itself, which collects and processes incoming signals from neighbouring cells. In this review, we suggest that the morphological and ionic properties of the dendrites, the soma and the axon provide these structures with an array of computational abilities that might enable them to contribute differentially to neuronal function.

Dendrites seem to be key players in functions such as binocular disparity ( Archie & Mel, 2000 ) and directional selectivity in the visual system of various species ( Single & Borst, 1998 ; Euler et al , 2002 ; Vaney & Taylor, 2002 ), as well as in improving sound localization ( Agmon-Snir et al , 1998 ) and in supporting the transition between encoding and retrieval modes of associative memory systems ( Hasselmo et al , 1996 ; Dvorak-Carbone & Schuman, 1999 ). Conversely, persistent activity maintained by somatic mechanisms has been suggested to represent a cellular correlate of working memory functions ( Goldman-Rakic, 1995 ). Finally, propagation delays of the action potential along the axon have been attributed a role in precise temporal coding in the auditory system of the barn owl ( Carr et al , 2001 ), whereas axonal Na + channels have been suggested to act as a memory reservoir for previous activity levels ( Segev & Schneidman, 1999 ).

Linking computational properties to behaviour is the ultimate challenge for both modelling and experimental studies of the future. Recent papers applying modelling, physiological, molecular, genetic and behavioural techniques in Drosophila and mice have shown the contribution of different voltage-dependent K + conductances in light processing by photoreceptors and in reversing age-induced impairments in learning and memory, respectively ( Vahasoyrinki et al , 2006 ; Murphy et al , 2004 ). Such multidisciplinary approaches—in which models are used to formulate experimentally testable predictions and experiments are used to verify the predictions and refine the models—will enable a more thorough investigation of how different neuronal components and the cell as a whole contribute to information processing capacity and behaviour.

The following open questions could provide fertile ground for collaborations among molecular biologists, geneticists, physiologists, modellers and behaviourists for further explorations of the mysteries of the brain. Do specific behaviours require certain neuronal computational tasks? Which parts of the neural circuit or the neuron itself are responsible for these tasks? What are the underlying molecular mechanisms for the distinct operating modes of neuronal integration? Such holistic approaches should lend support to the growing idea reinforced by this review: that something smaller than the cell lies at the heart of neural computation.

An external file that holds a picture, illustration, etc.
Object name is 7400789-i1.jpg

Eleftheria Kyriaki Pissadaki, Panayiota Poirazi (who is an EMBO Young Investigator) & Kyriaki Sidiropoulou

Acknowledgments

This work was supported by Alexander S. Onassis Public Benefit Foundation (K.S.), General Secretariat of Research and Technology, ΠENEΔ 01EΔ311 (E.K.P.) and the EMBO Young investigator Programme.

  • Agmon-Snir H, Carr CE, Rinzel J (1998) The role of dendrites in auditory coincidence detection . Nature 393 : 268–272 [ PubMed ] [ Google Scholar ]
  • Archie KA, Mel BW (2000) A model for intradendritic computation of binocular disparity . Nat Neurosci 3 : 54–63 [ PubMed ] [ Google Scholar ]
  • Ariav G, Polsky A, Schiller J (2003) Submillisecond precision of the input–output transformation function mediated by fast sodium dendritic spikes in basal dendrites of CA1 pyramidal neurons . J Neurosci 23 : 7750–7758 [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Azouz R (2005) Dynamic spatiotemporal synaptic integration in cortical neurons: neuronal gain, revisited . J Neurophysiol 94 : 2785–2796 [ PubMed ] [ Google Scholar ]
  • Baer SM, Rinzel J (1991) Propagation of dendritic spikes mediated by excitable spines: a continuum theory . J Neurophysiol 65 : 874–890 [ PubMed ] [ Google Scholar ]
  • Bernander O, Douglas RJ, Martin KA, Koch C (1991) Synaptic background activity influences spatiotemporal integration in single pyramidal cells . Proc Natl Acad Sci USA 88 : 11569–11573 [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Carr CE, Soares D, Parameshwaran S, Perney T (2001) Evolution and development of time coding systems . Curr Opin Neurobiol 11 : 727–733 [ PubMed ] [ Google Scholar ]
  • Cash S, Yuste R (1999) Linear summation of excitatory inputs by CA1 pyramidal neurons . Neuron 22 : 383–394 [ PubMed ] [ Google Scholar ]
  • Day M, Carr DB, Ulrich S, Ilijic E, Tkatch T, Surmeier DJ (2005) Dendritic excitability of mouse frontal cortex pyramidal neurons is shaped by the interaction among HCN, Kir2, and Kleak channels . J Neurosci 25 : 8776–8787 [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Debanne D (2004) Information processing in the axon . Nat Rev Neurosci 5 : 304–316 [ PubMed ] [ Google Scholar ]
  • Debanne D, Guerineau NC, Gahwiler BH, Thompson SM (1997) Action-potential propagation gated by an axonal I (A)-like K + conductance in hippocampus . Nature 389 : 286–289 [ PubMed ] [ Google Scholar ]
  • Dvorak-Carbone H, Schuman EM (1999) Patterned activity in stratum lacunosum moleculare inhibits CA1 pyramidal neuron firing . J Neurophysiol 82 : 3213–3222 [ PubMed ] [ Google Scholar ]
  • Egorov AV, Hamam BN, Fransen E, Hasselmo ME, Alonso AA (2002) Graded persistent activity in entorhinal cortex neurons . Nature 420 : 173–178 [ PubMed ] [ Google Scholar ]
  • Euler T, Detwiler PB, Denk W (2002) Directionally selective calcium signals in dendrites of starburst amacrine cells . Nature 418 : 845–852 [ PubMed ] [ Google Scholar ]
  • Gasparini S, Magee JC (2006) State-dependent dendritic computation in hippocampal CA1 pyramidal neurons . J Neurosci 26 : 2088–2100 [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Golding NL, Spruston N (1998) Dendritic sodium spikes are variable triggers of axonal action potentials in hippocampal CA1 pyramidal neurons . Neuron 21 : 1189–1200 [ PubMed ] [ Google Scholar ]
  • Golding NL, Mickus TJ, Katz Y, Kath WL, Spruston N (2005) Factors mediating powerful voltage attenuation along CA1 pyramidal neuron dendrites . J Physiol 568 : 69–82 [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Goldman-Rakic PS (1995) Cellular basis of working memory . Neuron 14 : 477–485 [ PubMed ] [ Google Scholar ]
  • Goldstein SS, Rall W (1974) Changes of action potential shape and velocity for changing core conductor geometry . Biophys J 14 : 731–757 [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Hasselmo ME, Wyble BP, Wallenstein GV (1996) Encoding and retrieval of episodic memories: role of cholinergic and GABAergic modulation in the hippocampus . Hippocampus 6 : 693–708 [ PubMed ] [ Google Scholar ]
  • Hausser M, Mel B (2003) Dendrites: bug or feature? Curr Opin Neurobiol 13 : 372–383 [ PubMed ] [ Google Scholar ]
  • Jarsky T, Roxin A, Kath WL, Spruston N (2005) Conditional dendritic spike propagation following distal synaptic activation of hippocampal CA1 pyramidal neurons . Nat Neurosci 8 : 1667–1676 [ PubMed ] [ Google Scholar ]
  • Kandel ER, Spencer WA (1961) Electrophysiology of hippocampal neurons. II. After-potentials and repetitive firing . J Neurophysiol 24 : 243–259 [ PubMed ] [ Google Scholar ]
  • Koch C, Segev I (2000) The role of single neurons in information processing . Nat Neurosci 3 ( Suppl ): 1171–1177 [ PubMed ] [ Google Scholar ]
  • Kopysova IL, Debanne D (1998) Critical role of axonal A-type K + channels and axonal geometry in the gating of action potential propagation along CA3 pyramidal cell axons: a simulation study . J Neurosci 18 : 7436–7451 [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Larkum ME, Zhu JJ, Sakmann B (2001) Dendritic mechanisms underlying the coupling of the dendritic with the axonal action potential initiation zone of adult rat layer 5 pyramidal neurons . J Physiol 533 : 447–466 [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Larkum ME, Senn W, Luscher HR (2004) Top-down dendritic input increases the gain of layer 5 pyramidal neurons . Cereb Cortex 14 : 1059–1070 [ PubMed ] [ Google Scholar ]
  • Lisman JE, Fellous JM, Wang XJ (1998) A role for NMDA-receptor channels in working memory . Nat Neurosci 1 : 273–275 [ PubMed ] [ Google Scholar ]
  • Loewenstein Y, Sompolinsky H (2003) Temporal integration by calcium dynamics in a model neuron . Nat Neurosci 6 : 961–967 [ PubMed ] [ Google Scholar ]
  • London M, Hausser M (2005) Dendritic computation . Annu Rev Neurosci 28 : 503–532 [ PubMed ] [ Google Scholar ]
  • London M, Segev I (2001) Synaptic scaling in vitro and in vivo . Nat Neurosci 4 : 853–855 [ PubMed ] [ Google Scholar ]
  • Losonczy A, Magee JC (2006) Integrative properties of radial oblique dendrites in hippocampal CA1 pyramidal neurons . Neuron 50 : 291–307 [ PubMed ] [ Google Scholar ]
  • Magee JC, Cook EP (2000) Somatic EPSP amplitude is independent of synapse location in hippocampal pyramidal neurons . Nat Neurosci 3 : 895–903 [ PubMed ] [ Google Scholar ]
  • Major G, Tank D (2004) Persistent neural activity: prevalence and mechanisms . Curr Opin Neurobiol 14 : 675–684 [ PubMed ] [ Google Scholar ]
  • Manor Y, Koch C, Segev I (1991) Effect of geometrical irregularities on propagation delay in axonal trees . Biophys J 60 : 1424–1437 [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • McCulloch WS, Pitts W (1943) A logical calculus of the ideas immanent in nervous activity . Bull Math Biophys 5 : 115–133 [ PubMed ] [ Google Scholar ]
  • Migliore M (2003) On the integration of subthreshold inputs from perforant path and Schaffer collaterals in hippocampal CA1 pyramidal neurons . J Comput Neurosci 14 : 185–192 [ PubMed ] [ Google Scholar ]
  • Migliore M, Messineo L, Ferrante M (2004) Dendritic Ih selectively blocks temporal summation of unsynchronized sistal inputs in CA1 pyramidal neurons . J Comput Neurosci 16 : 5–13 [ PubMed ] [ Google Scholar ]
  • Murphy GG, Fedorov NB, Giese KP, Ohno M, Friedman E, Chen R, Silva AJ (2004) Increased neuronal excitability, synaptic plasticity and learning in aged Kvβ1.1 knockout mice . Curr Biol 14 : 1907–1915 [ PubMed ] [ Google Scholar ]
  • Poirazi P (2005) Role of persistent activity in learning and memory capacity of young and aged CA1 pyramidal neuron models . Program No. 612. 8, 2005Abstract Viewer/Itinerary planner, Washington, DC, Society for Neuroscience, 2005 [http://www.sfn.org]
  • Poirazi P, Mel BW (2001) Impact of active dendrites and structural plasticity on the memory capacity of neural tissue . Neuron 29 : 779–796 [ PubMed ] [ Google Scholar ]
  • Poirazi P, Brannon T, Mel BW (2003) Arithmetic of subthreshold synaptic summation in a model CA1 pyramidal cell . Neuron 37 : 977–987 [ PubMed ] [ Google Scholar ]
  • Poirazi P, Brannon T, Mel BW (2003) Pyramidal neuron as two-layer neural network . Neuron 37 : 989–999 [ PubMed ] [ Google Scholar ]
  • Polsky A, Mel BW, Schiller J (2004) Computational subunits in thin dendrites of pyramidal cells . Nat Neurosci 7 : 621–627 [ PubMed ] [ Google Scholar ]
  • Rall W (1959) Branching dendritic trees and motoneuron membrane resistivity . Exp Neurol 1 : 491–527 [ PubMed ] [ Google Scholar ]
  • Rall W, Shepherd GM (1968) Theoretical reconstruction of field potentials and dendrodendritic synaptic interactions in olfactory bulb . J Neurophysiol 31 : 884–915 [ PubMed ] [ Google Scholar ]
  • Rall W, Burke RE, Smith TG, Nelson PG, Frank K (1967) Dendritic location of synapses and possible mechanisms for the monosynaptic EPSP in motoneurons . J Neurophysiol 30 : 1169–1193 [ PubMed ] [ Google Scholar ]
  • Ramon y Cajal (1933) Histology . Baltimore, MD, USA: Wood [ Google Scholar ]
  • Rudolph M, Destexhe A (2003) A fast-conducting, stochastic integrative mode for neocortical neurons in vivo . J Neurosci 23 : 2466–2476 [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Schiller J, Major G, Koester HJ, Schiller Y (2000) NMDA spikes in basal dendrites of cortical pyramidal neurons . Nature 404 : 285–289 [ PubMed ] [ Google Scholar ]
  • Segal M (2002) Changing views of Cajal's neuron: the case of the dendritic spine . Prog Brain Res 136 : 101–107 [ PubMed ] [ Google Scholar ]
  • Segev I, Rall W (1998) Excitable dendrites and spines: earlier theoretical insights elucidate recent direct observations . Trends Neurosci 21 : 453–460 [ PubMed ] [ Google Scholar ]
  • Segev I, Schneidman E (1999) Axons as computing devices: basic insights gained from models . J Physiol Paris 93 : 263–270 [ PubMed ] [ Google Scholar ]
  • Shepherd GM, Brayton RK, Miller JP, Segev I, Rinzel J, Rall W (1985) Signal enhancement in distal cortical dendrites by means of interactions between active dendritic spines . Proc Natl Acad Sci USA 82 : 2192–2195 [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Single S, Borst A (1998) Dendritic integration and its role in computing image velocity . Science 281 : 1848–1850 [ PubMed ] [ Google Scholar ]
  • Softky WR, Koch C (1993) The highly irregular firing of cortical cells is inconsistent with temporal integration of random EPSPs . J Neurosci 13 : 334–350 [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Stuart G, Spruston N, Sakmann B, Hausser M (1997) Action potential initiation and backpropagation in neurons of the mammalian CNS . Trends Neurosci 20 : 125–131 [ PubMed ] [ Google Scholar ]
  • Tahvildari B, Alonso A (2005) Morphological and electrophysiological properties of lateral entorhinal cortex layers II and III principal neurons . J Comp Neurol 491 : 123–140 [ PubMed ] [ Google Scholar ]
  • Tamas G, Szabadics J, Somogyi P (2002) Cell type- and subcellular position-dependent summation of unitary postsynaptic potentials in neocortical neurons . J Neurosci 22 : 740–747 [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Tegner J, Compte A, Wang XJ (2002) The dynamical stability of reverberatory neural circuits . Biol Cybern 87 : 471–481 [ PubMed ] [ Google Scholar ]
  • Traub RD, Wong RK, Miles R, Michelson H (1991) A model of a CA3 hippocampal pyramidal neuron incorporating voltage-clamp data on intrinsic conductances . J Neurophysiol 66 : 635–650 [ PubMed ] [ Google Scholar ]
  • Tsay D, Yuste R (2002) Role of dendritic spines in action potential backpropagation: a numerical simulation study . J Neurophysiol 88 : 2834–2845 [ PubMed ] [ Google Scholar ]
  • Tsay D, Yuste R (2004) On the electrical function of dendritic spines . Trends Neurosci 27 : 77–83 [ PubMed ] [ Google Scholar ]
  • Vahasoyrinki M, Niven JE, Hardie RC, Weckstrom M, Juusola M (2006) Robustness of neural coding in Drosophila photoreceptors in the absence of slow delayed rectifier K + channels . J Neurosci 26 : 2652–2660 [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Vaney DI, Taylor WR (2002) Direction selectivity in the retina . Curr Opin Neurobiol 12 : 405–410 [ PubMed ] [ Google Scholar ]
  • Wei DS, Mei YA, Bagal A, Kao JP, Thompson SM, Tang CM (2001) Compartmentalized and binary behavior of terminal dendrites in hippocampal pyramidal neurons . Science 293 : 2272–2275 [ PubMed ] [ Google Scholar ]
  • Wirth S, Yanike M, Frank LM, Smith AC, Brown EN, Suzuki WA (2003) Single neurons in the monkey hippocampus and learning of new associations . Science 300 : 1578–1581 [ PubMed ] [ Google Scholar ]
  • Yuste R, Urban R (2004) Dendritic spines and linear networks . J Physiol Paris 98 : 479–486 [ PubMed ] [ Google Scholar ]
  • Yuste R, Gutnick MJ, Saar D, Delaney KR, Tank DW (1994) Ca 2+ accumulations in dendrites of neocortical pyramidal neurons: an apical band and evidence for two functional compartments . Neuron 13 : 23–43 [ PubMed ] [ Google Scholar ]
  • Zhu JJ (2000) Maturation of layer 5 neocortical pyramidal neurons: amplifying salient layer 1 and layer 4 inputs by Ca 2+ action potentials in adult rat tuft dendrites . J Physiol 526 : 571–587 [ PMC free article ] [ PubMed ] [ Google Scholar ]

To revisit this article, visit My Profile, then View saved stories .

  • The Big Story
  • Newsletters
  • Steven Levy's Plaintext Column
  • WIRED Classics from the Archive
  • WIRED Insider
  • WIRED Consulting

This Is the Most Detailed Map of Brain Connections Ever Made

Image may contain Outdoors and Plastic

This image could be hung in a gallery, but it started life as a tiny chunk of a woman’s brain. In 2014, a woman undergoing surgery for epilepsy had a tiny chunk of her cerebral cortex removed. This cubic millimeter of tissue has allowed Harvard and Google researchers to produce the most detailed wiring diagram of the human brain that the world has ever seen.

Biologists and machine-learning experts spent 10 years building an interactive map of the brain tissue, which contains approximately 57,000 cells and 150 million synapses. It shows cells that wrap around themselves, pairs of cells that seem mirrored, and egg-shaped “objects” that, according to the research, defy categorization. This mind-blowingly complex diagram is expected to help drive forward scientific research, from understanding human neural circuits to potential treatments for disorders.

“If we map things at a very high resolution, see all the connections between different neurons, and analyze that at a large scale, we may be able to identify rules of wiring,” says Daniel Berger, one of the project’s lead researchers and a specialist in connectomics, which is the science of how individual neurons link to form functional networks. “From this, we may be able to make models that mechanistically explain how thinking works or memory is stored.”

Jeff Lichtman, a professor in molecular and cellular biology at Harvard, explains that researchers in his lab, led by Alex Shapson-Coe, created the brain map by taking subcellular pictures of the tissue using electron microscopy. The tissue from the 45-year-old woman’s brain was stained with heavy metals, which bind to lipid membranes in cells. This was done so that cells would be visible when viewed through an electron microscope, as heavy metals reflect electrons.

The tissue was then embedded in resin so that it could be cut into really thin slices, just 34 nanometers thick (in comparison, the thickness of a typical piece of paper is around 100,000 nanometers). This was done to make the mapping easier, says Berger—to transform a 3D problem into a 2D problem. After this, the team took electron microscope images of each 2D slice, which amounted to a mammoth 1.4 petabytes of data.

Once the Harvard researchers had these images, they did what many of us do when faced with a problem: They turned to Google. A team at the tech giant led by Viren Jain aligned the 2D images using machine-learning algorithms to produce 3D reconstructions with automatic segmentation, which is where components within an image—for example, different cell types—are automatically differentiated and categorized. Some of the segmentation required what Lichtman called “ground-truth data,” which involved Berger (who worked closely with Google’s team) manually redrawing some of the tissue by hand to further inform the algorithms.

Digital technology, Berger explains, enabled him to see all the cells in this tissue sample and color them differently depending on their size. Traditional methods of imaging neurons, such as coloring samples with a chemical known as the Golgi stain, which has been used for over a century, leave some elements of nervous tissue hidden.

Step Away From Screens With the 32 Best Family Board Games

In the example above, Berger made the smallest cells blue and the biggest cells red, with all other cells between falling on a color spectrum. This helped researchers to identify the brain’s six cortical layers and white matter.

While researchers have been able to identify structures from the data, one ongoing difficulty of the project is proofreading the automatic segmentation. This involves individuals manually sifting through every part of the 3D map to check for segmentation errors. “This is a huge challenge for human beings, because now we’re generating datasets that are larger than a single human can experience,” says Lichtman.

In parts of the data that have been proofread, Berger says that particular cells seem “really interested in contacting.” The researchers have found examples of over 50 synapses to one singular neuron, which, according to Berger, is a phenomenon previously overlooked that could be integral to cortical processing.

On top of identifying structures and connections, researchers have identified abnormal cells. Berger said he came across an unidentifiable egg-shaped “object” (much smaller than a cell body but part of a cell) when attempting to systematically categorize each cell in the dataset. Other ambiguous cells include those seemingly mirrored in shape and “tangled” cells that wrap around themselves; until further research is done, these cells remain mysteries. However, they may not remain so for long.

The brain map has been made open access, which means that these images have opened up boundless possibilities for progress in neuroscience, particularly as this is the first publicly available wiring diagram of the human brain at subcellular level. Both Berger and Lichtman emphasized that they did not go into the project with concrete aims of discovery but rather wanted to create the “possibility to observe,” and from this, they hope (and expect) that “further insights will come” from both the Lichtman lab and external researchers.

Berger anticipates that advancements could be made in understanding and treating mental conditions such as schizophrenia. Potential future discoveries could also expand beyond the mind, as Berger thinks the functions of the biological brain may be used to improve deep-learning AI systems and their structures.

In terms of future projects, the Harvard Lichtman lab plans to continue its collaboration with Google to “factor this rendering up another scale of a thousand” by studying a whole mouse brain. The research lab is also working on more human brain samples, to expand research into other regions of the brain. This will enhance the already invaluable resource and its ability to inform and expand future discoveries.

This article appears in the September/October 2024 issue of WIRED UK magazine.

You Might Also Like …

In your inbox: Our biggest stories , handpicked for you each day

How one bad CrowdStrike update crashed the world’s computers

The Big Story: How soon might the Atlantic Ocean break ?

Welcome to the internet's hyper-consumption era

research paper on brain anatomy

Grab your spot at the free arXiv Accessibility Forum

Help | Advanced Search

Computer Science > Computer Vision and Pattern Recognition

Title: imagen 3.

Abstract: We introduce Imagen 3, a latent diffusion model that generates high quality images from text prompts. We describe our quality and responsibility evaluations. Imagen 3 is preferred over other state-of-the-art (SOTA) models at the time of evaluation. In addition, we discuss issues around safety and representation, as well as methods we used to minimize the potential harm of our models.
Subjects: Computer Vision and Pattern Recognition (cs.CV)
Cite as: [cs.CV]
  (or [cs.CV] for this version)
  Focus to learn more arXiv-issued DOI via DataCite

Submission history

Access paper:.

  • Other Formats

license icon

References & Citations

  • Google Scholar
  • Semantic Scholar

BibTeX formatted citation

BibSonomy logo

Bibliographic and Citation Tools

Code, data and media associated with this article, recommenders and search tools.

  • Institution

arXivLabs: experimental projects with community collaborators

arXivLabs is a framework that allows collaborators to develop and share new arXiv features directly on our website.

Both individuals and organizations that work with arXivLabs have embraced and accepted our values of openness, community, excellence, and user data privacy. arXiv is committed to these values and only works with partners that adhere to them.

Have an idea for a project that will add value for arXiv's community? Learn more about arXivLabs .

IMAGES

  1. Anatomy of the Brain: [Essay Example], 586 words GradesFixer

    research paper on brain anatomy

  2. Anatomy And Physiology Of Brain Pdf » TestBookpdf.com

    research paper on brain anatomy

  3. Anatomy Human Brain Model Labeled

    research paper on brain anatomy

  4. [PDF] Secrets of the Brain: An Introduction to the Brain Anatomical

    research paper on brain anatomy

  5. (DOC) BRAIN ANATOMY

    research paper on brain anatomy

  6. Brain Anatomy

    research paper on brain anatomy

COMMENTS

  1. The anatomy of the brain

    In antiquity, the search for knowledge of the brain anatomy was marked by several protagonists who, over the centuries, produced proposals to justify the clinical findings of the period and define the functioning of the organ. The purpose of this article is to report, through temporal progress, who are the main characters who guided the path to ...

  2. What We Know About the Brain Structure-Function Relationship

    1. Introduction. How the human brain works is still an open question, as is its implication with brain architecture: the non-trivial structure-function relationship. The development of neuroimaging techniques, the improvement on their processing methods and the unfolding of computational neuroscience field have driven brain research focused ...

  3. Relating Structure and Function in the Human Brain: Relative ...

    Author Summary By analogy with the road network, the human brain is defined both by its anatomy (the 'roads'), that is, the way neurons are shaped, clustered together and connected to each others and its dynamics (the 'traffic'): electrical and chemical signals of various types, shapes and strength constantly propagate through the brain to support its sensorimotor and cognitive ...

  4. Unveiling universal aspects of the cellular anatomy of the brain

    We generate 2000 2D samples for each brain at each L value. In units of mip 4 voxels, we sample L = 2 n for integers 4 ≤ n ≤ 12 (up to L = 524.288 μm) in the human and mouse, and 4 ≤ n ≤ ...

  5. 2660 PDFs

    Explore the latest full-text research PDFs, articles, conference papers, preprints and more on BRAIN ANATOMY. Find methods information, sources, references or conduct a literature review on BRAIN ...

  6. How the human brain creates cognitive maps of related concepts

    The world is constantly changing. In a paper in Nature, Courellis et al. 1 explore, in impressive detail, how the human brain makes generalizations that enable it to adapt to change. As the ...

  7. Home

    Brain Structure & Function offers free color in print and online for all its papers!. Brain Structure & Function publishes research that provides insight into brain structure−function relationships. Studies published here integrate data spanning from molecular, cellular, developmental, and systems architecture to the neuroanatomy of behavior and cognitive functions.

  8. Geometric constraints on human brain function

    Abstract. The anatomy of the brain necessarily constrains its function, but precisely how remains unclear. The classical and dominant paradigm in neuroscience is that neuronal dynamics are driven ...

  9. Brain Structure and Function: the first 15 years—a retrospective

    History. The idea of a new journal, entitled Brain Structure and Function (BSAF), was conceived 15 years ago; in the Summer of 2006 in the office of Professor Karl Zilles, 1 then Chair of the Anatomy Department in Düsseldorf. Through visits in Düsseldorf and later in Jülich, where Dr. Zilles established a Neuroscience Research Institute in the former nuclear research facility, we had many ...

  10. An overview of the anatomy and physiology of the brain

    Even though the main objective of neuroscience research is to obtain a better knowledge of the brain, its parts and how its functions are connected to the mind and the nervous system of human beings, a majority of present efforts are aimed at answering minor issues with more comprehensive information about internal structures, physiology, and clinical implications in the central nervous system ...

  11. (PDF) The Brain and How it Functions

    The brain and its function. comprise a central nervous system. A brainstem, which in part relays information between the peripheral nerves and spinal. cord to the upper parts of the brain consists ...

  12. Revisiting the Functional Anatomy of the Human Brain: Toward a Meta

    Research inquiry, mainly using fMRI and the lesion method, has highlighted that mentalizing is performed by a brain-wide neural network extending to both the lateral and medial aspects on the brain, including the ventral precuneus/posterior cingulate cortex and the medial prefrontal cortex, the temporo-parietal junction along with the posterior ...

  13. HUMAN BRAIN ANATOMY

    Figure A-1 illustrates the major areas of the brain. The cerebrum is the general volume of the right and left hemisphere regions in the upper forebrain. It contains the cerebral cortex, a layer of cells, or gray matter, and the underlying connections, or white matter. The frontal lobe of the cerebral cortex is thought to be where important ...

  14. Brain Structure and Function: Insights from Chemical Neuroanatomy

    We present a brief historical and epistemological outline of investigations on the brain's structure and functions. These investigations have mainly been based on the intermingling of chemical anatomy, new techniques in the field of microscopy and computer-assisted morphometric methods. This intermingling has enabled extraordinary investigations to be carried out on brain circuits, leading ...

  15. Anatomy of the Brain

    The brain governs the functioning of all the body's other organs. It is also responsible for instinctual behaviors and integrates current information with past experiences. The brain makes us conscious, emotional, and intelligent. In humans, the brain is made up of about 100 billion nerve cells, called neurons; supporting cells (glia); and ...

  16. An Introduction to Human Brain Anatomy

    Abstract. This tutorial chapter provides an overview of the human brain anatomy. Knowledge of brain anatomy is fundamental to our understanding of cognitive processes in health and disease; moreover, anatomical constraints are vital for neurocomputational models and can be important for psychological theorizing as well.

  17. (PDF) Functional Anatomy of the Brain

    The central nervous system comprises the brain and spinal cord which provide sensation, control of movement, emotion, aesthetics, reason and self-awareness. The tissue that makes up the central ...

  18. Anatomical structures, cell types and biomarkers of the Human Reference

    The brain uses the 141 anatomical structures of the 'Allen Human Reference Atlas—3D, 2020' representing one half of the human brain 4; these structures were mirrored to arrive at a whole ...

  19. Anatomy, Central Nervous System

    Peripheral and Central Nervous Systems). The brain is an organ of nervous tissue responsible for responses, sensation, movement, emotions, communication, thought processing, and memory. The skull, meninges, and cerebrospinal fluids protect the human brain. The nervous tissue is extremely delicate and can be damaged by the smallest amount of force.

  20. Inside the brain of a neuron

    Introduction. Understanding how the brain works remains one of the most exciting and intricate challenges of modern biology. Despite the wealth of information that has accumulated during the past years about the molecular and biophysical mechanisms that underlie neuronal activity, similar advances have yet to be made in understanding the rules that govern information processing and the ...

  21. An Accurate and Rapidly Calibrating Speech Neuroprosthesis

    We refined the implantation targets using the Human Connectome Project multimodal MRI-derived cortical parcellation precisely mapped to the participant's brain 23 (Figure 1B and Fig. S2 and ...

  22. Brain Structure and Function: the first 15 years—a retrospective

    The idea of a new journal, entitled Brain Structure and Function (BSAF), was conceived 15 years ago; in the Summer of 2006 in the office of Professor Karl Zilles, Footnote 1 then Chair of the Anatomy Department in Düsseldorf. Through visits in Düsseldorf and later in Jülich, where Dr. Zilles established a Neuroscience Research Institute in the former nuclear research facility, we had many ...

  23. This Is the Most Detailed Map of Brain Connections Ever Made

    The tissue was then embedded in resin so that it could be cut into really thin slices, just 34 nanometers thick (in comparison, the thickness of a typical piece of paper is around 100,000 nanometers).

  24. An Introduction to Human Brain Anatomy

    The cerebellum is located under the cerebrum, its function is to coordinate muscle movements, maintain posture and balance. The brainstem includes the midbrain, pons, and medulla, its perform many ...

  25. Scientists Capture Clearest Glimpse of How Brain Cells Embody Thought

    Using electrical recordings from more than 3,000 neurons in 17 volunteers with epilepsy who were undergoing invasive monitoring in the hospital to locate the sources of their seizures, the researchers accrued a "uniquely revealing dataset that is letting us for the first time monitor how the brain's cells represent a learning process ...

  26. Research Snapshot: Drs. Marangelie Criado-Marrero, Sakthivel Ravi, and

    A new UF-led study sheds light on the interplay between aging and brain changes following repetitive mild traumatic brain injuries (TBI). The preclinical study, part of a line of research aimed at improving understanding of age-specific symptoms and risk factors associated with brain injury, was published July 30 in the journal NeuroImage.

  27. (PDF) Anatomy and Physiology of Brain in Context of ...

    Anatomy and Physiology of Brain in Context of Learning: A Review from Current Literature ... Conference Paper. Full-text available. ... Online Journal of Multidiciplinary Research 1(1): 2395-4892. ...

  28. Brain Machine Interface Gets a Map

    New article in Nature Comm on 'Column-based Brain-Machine Interface Accesses Higher Order Visual Circuits' doi: 10.1038/s41467-024-50375- ... Read the paper. ... One of the hottest research fields today is brain-machine interface (BMI), an area that integrates neuroscience, signal processing, and biosensor development. ...

  29. PDF Brain Structure and Function: the first 15 years a retrospective

    papers related to spinal cord due to increased amount of backlog. High quality anatomy research into the mammalian central nervous system is central to BSAF, although coverage may range beyond this taxon. 4 From business purpose Springer continued the numbering of the Anatomy and Embryology, a journal of which Zilles was Editor-in-Chief.

  30. [2408.07009] Imagen 3

    We introduce Imagen 3, a latent diffusion model that generates high quality images from text prompts. We describe our quality and responsibility evaluations. Imagen 3 is preferred over other state-of-the-art (SOTA) models at the time of evaluation. In addition, we discuss issues around safety and representation, as well as methods we used to minimize the potential harm of our models.