Abstract
Characterizing large-scale dynamic organization of the brain relies on both data-driven and mechanistic modeling, which demands a low versus high level of prior knowledge and assumptions about how constituents of the brain interact. However, the conceptual translation between the two is not straightforward. The present work aims to provide a bridge between data-driven and mechanistic modeling. We conceptualize brain dynamics as a complex landscape that is continuously modulated by internal and external changes. The modulation can induce transitions between one stable brain state (attractor) to another. Here, we provide a novel method—Temporal Mapper—built upon established tools from the field of topological data analysis to retrieve the network of attractor transitions from time series data alone. For theoretical validation, we use a biophysical network model to induce transitions in a controlled manner, which provides simulated time series equipped with a ground-truth attractor transition network. Our approach reconstructs the ground-truth transition network from simulated time series data better than existing time-varying approaches. For empirical relevance, we apply our approach to fMRI data gathered during a continuous multitask experiment. We found that occupancy of the high-degree nodes and cycles of the transition network was significantly associated with subjects’ behavioral performance. Taken together, we provide an important first step toward integrating data-driven and mechanistic modeling of brain dynamics.
Author Summary
Brain dynamics are often described by data-driven models or mechanistic dynamical systems models to understand how specific brain states persist or change (transition). However, there lacks a computational framework that explicitly connects states/transitions discovered by data-driven methods to those of mechanistic models, leading to a disconnection between data analysis and theoretical modeling. To begin bridging this gap, we develop a data-driven method, the Temporal Mapper, to extract dynamical systems features from time series and represent them as attractor transition networks. The Temporal Mapper can reconstruct ground-truth transition networks of mechanistic models. When applied to human fMRI data, the method helps predict behavioral performance from the topology of transition networks. Potential applications include characterizing brain dynamic organization in health and diseases and designing brain stimulation protocols.
INTRODUCTION
The brain exhibits complex dynamics (Buzsaki, 2006; Kelso, 1995). Characterizing its overall dynamic organization is a fundamental step in assessing brain functions and brain fingerprinting for healthy individuals and patients with psychiatric disorders (Saggar & Uddin, 2019). One common approach is to infer dominant “brain states” and the transitions between them from neuroimaging time series data (e.g., Cavanna et al., 2018; Li et al., 2017; Meer et al., 2020; Saggar et al., 2018; Taghia et al., 2018; Tang et al., 2012; Zalesky et al., 2014). Such “states” and transitions can be defined by a diverse array of data-driven methods. Here we categorized a model as data driven if it does not require additional knowledge of the brain other than the time series data recorded from it. On the other side of the spectrum, brain dynamics are often modeled by large-scale nonlinear dynamical systems models with various levels of biophysical details (Breakspear, 2017; Deco et al., 2011). Here we categorize this type of model as mechanistic, as they aim to describe the dynamical mechanism of interaction between constituents of the brain, which requires prior knowledge or assumptions about the biophysical and anatomical features of the brain in addition to the time series data measured. States and transitions discovered using data-driven methods often share conceptual appeal to nonlinear dynamics concepts such as attractors (stable states) and phase transitions. Yet, a direct link between data-driven and mechanistic modeling of the brain remains missing. In this work, we develop a data analysis method to represent time series data as a directed graph, whose nodes and edges could reasonably map directly to the underlying attractors and phase transitions in a nonlinear dynamic model of the brain. We first validate our method by using simulated transitions and then apply the method to human fMRI data to demonstrate its empirical relevance in assessing transitions associated with cognitive task switching. This work helps build the missing link between data-driven and mechanistic modeling of complex brain dynamics. With a direct link to mechanistic models, data-driven models may better inform experimenters and clinicians of the network effect of causal perturbation (e.g., transcranial magnetic stimulation) in basic neuroscience and in the treatment of psychiatric disorders.
A signature of nonlinear brain dynamics is multistability, that is, the coexistence of multiple stable brain activity patterns (Kelso, 2012), which may be referred to as attractors in technical terms or persistent brain states colloquially. Transitions between these brain states may occur either driven by external influences or internal dynamics. Intermittent visits to different brain states are often referred to as metastability (Tognoli & Kelso, 2014). Multistability and metastability—the existence of and the transitions between different brain states—are key elements in the mechanistic modeling of brain dynamics and functional connectivity (FC) (van den Heuvel & Hulshoff Pol, 2010). Typically, such modeling approaches use large-scale biophysical network models that also incorporate biologically informed parameters and the human structural connectome (Deco et al., 2011, 2013, 2014; Deco & Jirsa, 2012; Golos et al., 2015; Hansen et al., 2015; Zhang et al., 2022).
The mechanistic modeling of state transitions in large-scale brain dynamics was motivated by, among other things, the observations of how large-scale FC patterns vary as a function of time, that is, the dynamic functional connectivity (dFC) (Hutchison et al., 2013; Preti et al., 2017). dFC patterns are primarily computed as correlation coefficients between time series within a sliding window. More recently, single time-frame methods (e.g., Faskowitz et al., 2020; Esfahlani et al., 2020) have been developed to tackle FC analysis at the finest temporal resolution and reduce the amount of data needed for stably estimating dFC patterns (Laumann et al., 2017; Leonardi & Van De Ville, 2015). Altogether, (d)FC analyses play a central role in the empirical understanding of brain dynamic organization. Abnormal FC and abnormal transitions between dFC patterns have been linked to a wide spectrum of psychiatric and neurological disorders (Barber et al., 2018; Díez-Cirarda et al., 2018; Du et al., 2021; Fox & Greicius, 2010; Garrity et al., 2007; Lui et al., 2011; Rabany et al., 2019; Saggar & Uddin, 2019).
What remains unclear is to what extent dFC patterns can be mapped to dynamical systems concepts such as attractors or stable states. With a data-driven approach, dFC patterns that repeat in time can be assigned to a relatively small number of “FC states” using, for example, clustering methods (Allen et al., 2014) or hidden Markov models (Quinn et al., 2018; Rabiner, 1989; Vidaurre et al., 2017). However, directly conceptualizing FC states as dynamical system states or attractors is not easy, especially when one needs to write down the differential equations governing the state evolution. Thus, mechanistic models of large-scale brain dynamics typically use mean-field neural activity (Cabral et al., 2017) (e.g., population firing rate, the fraction of open synaptic channels) or its derived BOLD signal (Friston et al., 2003), rather than vectorized dFC patterns, as state variables. FC states can be derived post hoc from simulated neural dynamics (Golos et al., 2015; Hansen et al., 2015), but a direct correspondence between such post hoc FC states and dynamical system attractors is yet to be demonstrated. Our recent modeling work suggests that FC patterns may be signatures of phase transitions between stable states rather than the states themselves (Zhang et al., 2022). All the above point to the need for a data-driven method to quantify stable brain states and transitions directly from time series data and allow mapping of such states/transitions to underlying attractors and phase transitions derived from mechanistic modeling.
In the present work, we leverage existing methods of computational topology/geometry and large-scale biophysical network modeling to bridge this gap. Topological and geometrical analysis of dynamical systems traces back to the time of Poincaré (Poincaré, 1967). However, efficient computational infrastructure for generalizing such methods to higher dimensional dynamics was not in place until recently. Morse decomposition has been used in the rigorous analysis of nonlinear dynamical systems, for example, to represent a dynamical system as a directed graph whose nodes map to attractors (and repellers) and edges to transitions (connecting orbits) (Cummins et al., 2016; Kalies et al., 2005). However, neuroimaging data live in a very high dimensional space sparsely covered by samples, which renders many rigorous methods inapplicable. With a data-driven approach, combinatorial representations (e.g., graphs or simplicial complexes) of neural time series or FC patterns can be generated using existing topological data analysis (TDA) tools such as Mapper (Carlsson, 2009; Geniesse et al., 2019, 2022; Saggar et al., 2018, 2022; Singh et al., 2007) and persistent homology (Billings et al., 2021; Carlsson, 2009; Chazal & Michel, 2021; Edelsbrunner & Morozov, 2013; Giusti et al., 2015; Petri et al., 2014). In between, there are emerging efforts to develop dynamical system-oriented TDA methods (Garland et al., 2016; Kim & Mémoli, 2021; Munch, 2013; Myers et al., 2019; Perea, 2019; Tymochko et al., 2020), some specifically supported by mechanistic models of biological dynamics (Gameiro et al., 2004; Topaz et al., 2015; Ulmer et al., 2019; Zhang et al., 2020). The present work falls under this in-between category, building on our previous work on the TDA (Geniesse et al., 2019, 2022; Saggar et al., 2018) and biophysical network modeling of large-scale brain dynamics (Zhang et al., 2022).
In the current work, our contribution is threefold. First, we introduce a novel method to extract features associated with dynamical systems (i.e., attractors and their directed transitions) from the observed time series data alone. Second, to validate our approach, we develop a method to simulate a complex sequence of phase transitions in a large-scale neural dynamic model in a controlled manner. This simulated data not only provides a ground truth of the coexisting attractors in the model and their respective transitions, but also allows examination of intricate but relevant nonlinear dynamic concepts such as hysteresis. Third, we apply our method to a real-world human fMRI dataset to examine the efficacy of our method in capturing states and their transitions associated with cognitive task switching from time series data alone. Taken together, we provide a critical methodological step toward bridging the mechanistic and data-driven modeling of large-scale brain dynamics.
RESULTS
In this section, for larger accessibility of our results, we first provide a brief introduction to the key nonlinear dynamics concepts and intuitions. We then introduce the framework to simulate a complex sequence of phase transitions using a large-scale neural dynamic model. Finally, we present an introduction to our Temporal Mapper approach and its application to simulated as well as real-world fMRI datasets.
Nonlinear Dynamics and the Brain
Brain activities can be thought of as dynamics unfolding on a nonlinear landscape (Figure 1A). Each state in this landscape represents a pattern of activation over the whole brain. Some states are stable, which are termed attractors (Figure 1A, 1–3) since they attract nearby states to evolve toward them. The coexistence of multiple attractors—multistability—is a signature of the brain’s dynamic complexity (Kelso, 2012). The landscape can be shaped by a variety of intrinsic and extrinsic factors, such as external input, synaptic conductance, and structural connectivity (Zhang et al., 2022). Theoretically, these factors are often considered control parameters (Figure 1B).
As the control parameter changes, the landscape deforms with it. For illustration, sliding the gray plane in Figure 1B up and down the control parameter axis changes the landscape within the plane. With sufficient multistability, changing a control parameter back and forth along the same path (Figure 1C, dashed lines below the horizontal axis) can lead to distinct paths of transitions between attractors (Figure 1C, dashed lines in the bifurcation diagram above the horizontal axis)—a feature known as hysteresis. Due to the asymmetry in the path of transitions, directed graphs are better suited to minimally represent the transition network (Figure 1D), where the nodes map to the attractors visited (nodes 1–3) and edges map to the transitions between attractors.
The topological complexity of this attractor transition network reflects the brain’s intrinsic complexity through its interaction with the internal or external environment. In the present work, we develop a method to retrieve such transition networks from simulated neural dynamics and human fMRI data. In the following sections, we demonstrate that the networks reconstructed from simulated data are reasonable approximations of the theoretical ground truth, and those constructed from fMRI data help predict human behavioral performance.
Computational Framework to Simulate Complex Sequences of Phase Transitions and Represent Them as an Attractor Transition Network
In this subsection, we introduce the computational framework used to simulate neural dynamics. Simulations convey several advantages: (a) we can parametrically control and induce transitions between attractors, (b) we can compute the ground-truth transition network given the exact equations, and (c) we can directly compare the reconstructed network (from simulated time series alone without knowing the equations or any parameters) to the ground truth to assess the quality of reconstruction.
For simulations and computing the ground-truth transition network, we choose a biophysically informed model of human brain dynamics (Figure 2A; for details, see section Biophysical Network Model of the Human Brain in Methods). The model describes the dynamics of a network of 66 brain regions, shown to capture FC patterns in human resting fMRI (Zhang et al., 2022) (the cartoon in Figure 2A includes six regions only for illustrative purposes). The model is an adaptation of the reduced Wong–Wang model (Deco et al., 2014; Wong & Wang, 2006) in the form of the Wilson–Cowan model (Wilson & Cowan, 1972, 1973) with improved multistability (Zhang et al., 2022). Each region i consists of an excitatory population (E) and an inhibitory population (I), with associated state variables (SE(i), SI(i)). Long-range connections between regions (Cij in Figure 2A) are defined by the human connectome using data from the Human Connectome Project (Civier et al., 2019; Van Essen et al., 2013) (see Methods for details). The overall strength of global interaction is modulated by an additional global coupling parameter G. We define G as the control parameter, whose dynamics (Figure 2B) modulate the shape of the underlying dynamic landscape and induce transitions between attractors through bifurcation (see bifurcation diagram Figure 2D). The simulated neural dynamics in this time-varying landscape are shown in Figure 2C. It is important to note that here we assume the control parameter G, and consequently the shape of the underlying landscape itself, is changing much slower than the state dynamics occurring within the landscape (the ball in Figure 1A can roll quickly into the valley when the landscape has barely deformed). In other words, the present conceptual framework assumes a separation of time scale between the dynamics of the control parameter (e.g., G) and intrinsic state dynamics (e.g., defined in Equations 1 and 2 by the time constants τE and τI for the excitatory and inhibitory neuronal population, respectively). Physiologically, the changes in global coupling G can be interpreted as changes in the arousal level due to, for example, task demands. Recent work of Munn and colleagues (2021) suggests that cortical dynamic landscapes are modulated by ascending subcortical arousal systems mediated by the locus coeruleus (adrenergic) and the basal nucleus of Meynert (cholinergic). In particular, the locus coeruleus-mediated system promotes global integration across the cortex and reduces the energy barrier for state transitions.
Our methodological goal is to recover the cross-attractor transitions from the simulated neural dynamics (the gating variables SE(i)) and the BOLD signals derived from them (down-sampled to TR = 720 ms as in the Human Connectome Project (Van Essen et al., 2013)). The transitions can be encapsulated as a transition network (Figure 2E) and unfolded in time as a recurrence plot (Figure 2F). The recurrence plot depicts how far away the attractor occupied at each time point is from that of every other time point. Here, “how far away” is measured by the shortest path length from one node to another in the attractor transition network instead of the Euclidean distance between states in the original state space. The path length takes into account the underlying dynamics: two states can be far away in the state space but closely connected by transitions in the dynamical system, and conversely, two states can be close in the state space, but it could be costly to transition between each other against the underlying dynamics. The theoretical ground truth (Figure 2E and F) is constructed by assigning each time point to an attractor (Figure 2D and E) precomputed from Equations 4 and 5 (see section Computation of the Attractor Repertoire and Ground-Truth Transition Network in Methods). Computation of the ground truth requires all model parameters, including the state variables SE(i), SI(i), and the control parameter G, for example. As depicted, the transition network is directed to capture the “flow” of time. Further, the size of the node in the transition network represents how many sequential time points map onto that attractor, that is, the dwell time.
We assess the efficacy of our and others’ methods by comparing the reconstructed networks (Figure 3Bi and Bii) and recurrence plots (Figure 3Ci and Cii) to the ground truth (Figure 2E and F).
Temporal Mapper to Reconstruct Attractor Transition Network From Time Series Alone
To reconstruct the attractor transition network, using only time series data, our Temporal Mapper approach first constructs a temporal version of the k nearest neighbor (kNN) graph from samples in the time series (Figure 3Ai–iv). The time series data from multiple brain regions (Figure 3Ai) are first used to compute the pairwise distance between time points in Euclidean space (Figure 3Aii). This distance matrix is used to determine the k-nearest neighbors for each time point. Time points that are reciprocal k-nearest neighbors (excluding temporal neighbors) are connected by edges, forming the spatial kNN graph (Figure 3Aiii). Reciprocal edges in the neighborhood graph (Figure 3Aiii) connect spatial neighbors, while directed edges connect temporal neighbors, indicating the “arrow of time” (Figure 3Aiv). Nodes that are close to each other in the neighborhood graph are contracted to a single node in its compressed version (Figure 3Av). We consider the compressed graphs (Figure 3Av) as reconstructed attractor transition networks. Further details of this construction are provided in section Temporal Mapper–Construction of Transition Networks in Methods. Visually, the reconstructions (Figure 3Bi and Ci) are reasonable approximations of the ground truth (Figure 2E and F). This result is confirmed by quantitative comparisons against permuted time series (Figure 3D and E) and phase-randomized time series (Figure S1). The reconstruction remains robust at lower sampling rates (e.g., TR = 2 s; see Supporting Information Figures S11 and S12 for reconstruction accuracy drop-off rate under down-sampling) and across different realizations and integration time steps (Figure S14). See section Gromov–Wasserstein Distance Between Transition Networks in Methods for details of the graph dissimilarity measures used in these comparisons.
The transition network, BOLD signals, and dFC reveal different facets of brain dynamics.
Next, we use the simulated data to compare the dynamics in the reconstructed transition networks to its associated BOLD dynamics (from which the network is reconstructed) and dFC. Given the a priori knowledge of the ground truth, we are able to examine how different representations of the simulated time series capture different aspects of the model brain dynamics.
Dynamics in different spaces of representation (e.g., transition network, BOLD, dFC) can be compared in terms of the recurrence plots—a distance matrix that describes how far away the states at any two time points are from each other. The recurrence plot of the ground-truth transition network (Figure 4A, reproduced from Figure 2G) and that of the control parameter G (Figure 4B) provide an a priori reference for comparing the reconstructed network (Figure 4C), BOLD (Figure 4D), and dFC (Figure 4E). In the networks (Figure 4A and C), the interstate distance used to compute the recurrence plots is the shortest path length from one node to another. For the control parameter G and BOLD (Figure 4B and D), the distance is simply the Euclidean distance between states. For dFC (Figure 4E), the distance is the Euclidean distance between vectorized dFC matrices (Fisher-z transformed elements in the lower triangle) computed in a 30-TR sliding window. fMRI data in 20–30 TR windows have been shown to generate stable correlations (Allen et al., 2014; Hutchison et al., 2013).
Dynamics in the ground truth and the reconstructed transition networks can also be represented as the state (attractor) dynamics shown in Figure 4G and H, respectively. In this representation, we can visualize which attractors are visited during the course of time. Here, we sorted (and colored) the attractors based on their SE values (i.e., higher values mean more excitation). As evident, the dynamics extracted using the reconstructed transition network (from the Temporal Mapper approach) closely follow the dynamics of the ground-truth network, indicating the nodes of the Temporal Mapper network map onto underlying dynamical attractors reasonably well. A similar visualization for BOLD and dFC can be constructed using traditional community detection approaches on the recurrence matrix (see Figure S3 for more details).
The recurrence plot for the ground-truth transition network is more complex than that of the control parameter G (Figure 4A vs. B)—this added complexity reflects that of the underlying dynamic landscape, which is what we aim to examine. In terms of the state dynamics, the model brain passed through a sequence of attractors and transitions in a path-dependent manner (gray areas, Figure 4G), while the control parameter undergoes continuous and reversible changes (Figure 4F, color coded by attractor index to show the model brain can visit distinct attractors given the same level of G). Such path dependency in the ground-truth transition network is indicative of multistability and hysteresis of the underlying dynamic landscape (cf. Figure 1C). Thus, the discrete sequence of attractor transitions (gray areas, Figure 4G) is the signature of the model brain dynamics. The reconstructed transition network (Figure 4C) reasonably approximates the ground truths (Figure 4A) both in terms of the recurrence plot (Figure 4C) and the state dynamics (Figure 4H). Though some detailed transitions in the ground truth are not resolved by the reconstructed network, it is not surprising due to the low-pass filtering effect of the hemodynamic response—faster neural activity may not be recoverable from BOLD signals in principle. The recurrence plot of the simulated BOLD (Figure 4D) to a large extent follows the control parameter G, though some transitions are already visible without further state detection. Interestingly, the recurrence plot of dFC (Figure 4E) approximates neither that of the ground-truth transition network nor that of the parameter G. dFC does not differentiate distinct attractors and exhibits a mixed reaction to transitions and parameter changes (see Figure S3D for how dFC states differ from attractors states in Figure S3B and BOLD states in Figure S3C).
A further comparison between the reconstructed transition network, BOLD, and dFC is shown in Figure 4I as the row (or column) averages of the corresponding recurrence plots, that is, the average distance from the state/attractor at each time point to all other time points. This simplified representation helps us visualize the difference between dynamics in different spaces. This method was used by Saggar et al. (2018) to examine transitions between tasks in the aforementioned continuous multitask experiment (Gonzalez-Castillo et al., 2015). To understand what information each representation captures, we separate the dynamics into two different regimes: one with a single highly persistent attractor (white areas in Figure 4G), and one with a sequence of transitions between less persistent attractors (gray area). The average distance in the reconstructed network (blue, purple curves in Figure 4I) and BOLD (red) change little in the single-attractor regime (white areas). In the same single-attractor regime, the average distance of dFC (yellow curve in Figure 4I) appears to track the time derivative of the control parameter G only when G is decreasing (2nd and 4th white areas). Clearly, dFC states are not equivalent to persistent attractors. The distinction between the single-attractor and the transition regimes (white vs. gray areas in Figure 4I) is best indicated by the average distance in the reconstructed network (blue, purple curve). Specifically, the persistent attractors are more source-like (high sink distance, low source distance—purple above blue curve in the white areas of Figure 4I) while the sequence of transitions between less persistent attractors are more sink like (high source distance, low sink distance—purple below blue curve in the gray areas of Figure 4I). In contrast, the average distance of BOLD (Figure 4I, red) best mirrors the value of G (Figure 4F high G is associated with low average BOLD distance and vice versa). In short, the dynamics in the reconstructed transition network most effectively separate the regimes of attractor persistence and transitions in the model brain, compared to BOLD and dFC.
Application of Temporal Mapper to Real Human fMRI Dataset
Following the above theoretical analysis, we now apply the method to real human fMRI data to characterize the dynamic landscape of the human brain. We examine what features of the reconstructed transition networks are relevant to cognitive performance to demonstrate how the method developed above can be used in empirical settings. Data from 18 subjects were acquired from a continuous multitask experiment (Gonzalez-Castillo et al., 2015). Each subject performed eight blocks (color coded in Figure 5G) of tasks (memory, video, math) or rest in a single 25.4-min scan. The theoretical model used in the previous sections was designed to reflect this type of block design using the time-varying parameter G. During construction of the transition networks, we set the parameters to k = 5 and δ = 2. Justification for this choice as well as further parameter perturbation analysis is reported in Figures S9 and S10.
Figure 5A and D show the reconstructed transition network and the corresponding recurrence plot for a subject with good behavioral performance (top four in accuracy and reaction time), and Figure 5B and E for a subject with bad behavioral performance (bottom four in accuracy and reaction time). For both subjects, the central nodes are occupied mainly during the memory and math tasks (yellow, red nodes in Figure 5A and B) and go on excursions during the video task and rest (orange, green nodes). In aggregate across all subjects (Figure 5C), the memory task clearly dominates the highest degree nodes (yellow bars on the left end). In contrast, rest and the video task dominate the long excursions through the highest degree nodes (green and orange bars on the right end in Figure 5F). Later, we use this separation between the more cognitively demanding tasks (memory and math) and the relatively less demanding tasks (video and rest) over different types of nodes to predict subjects’ behavior performance (Figure 6).
Similar to Figure 4I for the theoretical model, here we examine the dynamics in the transition networks constructed from human fMRI data in Figure 5G as the row average (source distance, blue) and column average (sink distance, red) of the recurrence plots. Both the source and sink distance clearly track the transitions between blocks. Interestingly, both math and memory are associated with a low average distance in contrast to video and rest, which suggests brain states associated with math and memory are in the more central part of the transition network. The observation is consistent with our previous findings using related methods, Mapper (Saggar et al., 2018) and NeuMapper (Geniesse et al., 2022), where the task-positive brain activation patterns were found concentrated at the core of the graph with a clear core-periphery structure. Figure 5H shows the difference between the source and sink distance. When this sink-source difference deviates from zero (black dashed line), the brain occupies a node that is either more of a source to other nodes (source-like, negative values) or more of a sink to other nodes (sink-like, positive values; see Figure S6 for example networks). The source-like or sink-like deviation is visibly more prominent near the transitions between tasks, for example, block 2 in Figure 5H. This observation is verified statistically in Figure S7A. The source-like or sink-like deviation is also more prominent during rest and the video task than during the memory and math tasks (cf. Figure S7B). A closer examination of the dynamics in Figure 5H reveals that the brain tends to enter the memory and math tasks through source-like nodes (downward arrow in Figure 5H) and exit the block through sink-like nodes (upward arrow in Figure 5H). This is not the case for rest and the video task. The corresponding statistical verification is shown in Figure S8. This may be due to the fact that math and memory tasks are more structured such that the brain enters the task via specific source states, while the brain can enter the resting state, for example, from any state. In short, the transition networks and recurrence plots exhibit multiple features that keep track of the task and block structure.
Figure 6 shows how features of the transition networks relate to subjects’ behavioral performance. Greater separation between cognitively demanding tasks (memory and math) and less demanding tasks (video and rest) over the highest degree nodes predicts a higher percentage of correct responses (Figure 6A), a faster reaction time (Figure 6B), and fewer missed trials (Figure 6C). The statistical significance of the correlation coefficients is further validated against 1,000 random permutations of the subject-performance correspondence (95% confidence intervals: −0.4697, 0.5163; −0.5123, 0.4302; and −0.5340, 0.4342 for Figure 6A, B, and C, respectively). The number of cycles connecting each high-degree node back to itself also exhibits behavioral relevance (Figure 6D–F). On average, a greater number of cycles is strongly associated with slower reaction time (F(1, 400) = 46.63, p < 10−10; Figure 6E). There is also a scale-specific effect—cycles around length 10 are especially predictive of slower reactions. Here the cycle length can be roughly interpreted as the number of TRs required to leave and return to a high-degree node. To a lesser extent, a greater total number of cycles is also associated with a greater percentage of correct responses (F(1, 400) = 4.17, p = 0.014). There is no statistically significant relation between the number of cycles and the percentage of missed trials. In short, high-degree nodes and the excursions from them, that is, cycles, are key behavioral relevant features.
DISCUSSION
In the present work, we propose a computational method for reconstructing attractor transition networks from neural time series, named the Temporal Mapper. The method represents a time series as a directed graph whose nodes and edges map to attractors and phase transitions in the underlying dynamical system. In particular, the method addresses the scenario where the underlying system is nonstationary or nonautonomous, as, for example, when the brain is under continuously varying task demand, environmental changes, arousal levels. Using simulated brain dynamics, we demonstrate that the method provides a good approximation of the theoretical ground truth of cross-attractor transitions. Applying the method to human fMRI data, we show that the dynamics in the reconstructed networks clearly track the transitions between tasks. High-degree nodes and cycles of the network are key features that help predict human behavioral performance. Together, the theoretical and empirical analyses provide a basic theoretical and computational framework for bridging the data-driven and mechanistic modeling of brain states and transitions.
The present method builds on our previous works on characterizing neural dynamic landscape using TDA. In particular, it belongs to a family of neural time series analysis tools (Geniesse et al., 2019, 2022; Saggar et al., 2018, 2022) based on Mapper (Singh et al., 2007). NeuMapper (Geniesse et al., 2022) is the closest methodological match to Temporal Mapper in that it uses a reciprocal kNN graph construction without using local low-dimensional embeddings as an intermediate step. In another variant of Mapper approach, used in Saggar et al. (2022), an embedding step is typically utilized to examine the latent space, with any chosen dimension reduction techniques. Across the family of Mapper-based tools and related topological methods (Carlsson, 2009; Munch, 2013), time series are commonly treated as distributions of sample points in the state space. Constructing a topological representation, for example, a graph, from such a distribution often concerns only the spatial distance between points in the state space or that of their lower dimensional projections. The fact that these sample points are part of a time series—that there is an explicit sequential relation between sample points—is unutilized. Specifically, within the realm of fMRI data, which has been traditionally studied using time series techniques, previous applications of Mapper (Geniesse et al., 2022; Saggar et al., 2018, 2022) have focused on the geometric distribution of sample points and discarded the temporal information in the sequence of sample points. The present method is designed precisely to take advantage of this sequential information, that is, the arrow of time. Incorporating the arrow of time better reflects the fact that the system that generates the time series, for example, the brain, is a dynamical system. That is, the subsequent states of the system depend on the current state, and the exact nature of this dependency is the nature of the dynamical system. Restoring this temporal dependency in the construction has several consequences that we describe next.
First, the arrow of time restores the connectivity between sample points that could be far apart in the state space but tightly linked dynamically. The implication is most significant during a phase transition (vertical dashed lines in Figure 1C). At a transition, the dynamical system jumps suddenly from one attractor to another at a time scale much faster than the state dynamics within the same attractor. The combination of high velocity and even sampling in time makes consecutive sample points far apart in the state space, despite the direct dynamic dependency between them. Without the arrow of time, the increase of velocity during transitions is unaccounted for. The spatial information alone cannot determine the dynamic linkage between states of the system during the transitions, which happen to be key moments of interest.
Second, path lengths in the transition networks carry dynamical rather than purely geometrical information. The arrow of time introduces directedness into the networks. Thus, the shortest path lengths between two nodes are no longer symmetric and thus cannot be interpreted as geometric distance. The arrow of time attaches information about the underlying dynamic landscape to the state space. At an intuitive level, the shortest path from node x to node y can be considered the path of least time, or least action, within the landscape. In other words, paths in the transition networks could putatively encode actions in the state space.
Lastly, incorporating the arrow of time distinguishes the present method from common point cloud–based TDA techniques. Point cloud data—sets of disconnected points—are fundamental substrates of topological or geometrical analysis and learning. Such analysis includes nonlinear dimension reduction techniques such as Isomap (Tenenbaum et al., 2000) or Laplacian Eigenmaps (Belkin & Niyogi, 2003), TDA methodologies such as persistent homology (Carlsson, 2009; Edelsbrunner & Morozov, 2013), and deep learning (Qi et al., 2017). With the points first connected to their temporal neighbors, the present method operates on, in essence, a discretized version of a curve with a direction defined on it—naturally depicting a trajectory of an autonomous, deterministic dynamical system. Constructing a spatiotemporal neighborhood graph (Figure 3B) is thus folding a directed curve rather than “connecting dots.” An exposition of the mathematical consequences of the construction is beyond the scope of the present work but worthy of further study, especially with regard to its behavior as the sampling rate of the time series approaches infinity and when multiple trajectories are included in the construction.
One may find the transition networks in the present work reminiscent of hidden Markov models (HMM) for detecting discrete states in brain dynamics and transition probabilities between states (Baker et al., 2014; Meer et al., 2020; Ou et al., 2013; Rezek & Roberts, 2005; Vidaurre et al., 2017). A key distinction is that the number of discrete states in an HMM is set a priori, while its analog in the present method—the number of attractors visited—is data driven. The dynamic landscape of the human brain can be highly complex and sensitive to the organization of the large-scale structural connectome (Zhang et al., 2022). There is no a priori way to determine how many attractors may be visited during tasks and rest. Moreover, the dynamic landscape of each subject is shaped by the structural connectivity in each subject’s own brain. It cannot be assumed that the same number of attractors would be visited across subjects. Thus, the present method presents a flexible framework that requires fewer assumptions about the underlying dynamical system. For example, the Markov property for HMM (dependency on present state only) may not be satisfied in nonautonomous dynamical systems as conceptualized in the present work (see Figure S13 for an application of HMM on the simulated data), while the Temporal Mapper does not require the Markov property. Apart from statistical inference methods such as the HMM, topological representations of attractor networks also figure in differential equation-oriented state space decomposition methods, for example, based on the Conley Index Theory (Ban & Kalies, 2006; Kalies et al., 2005). In comparison to such mathematically rigorous approaches, the present method better accommodates empirical applications where the state space is often sparsely covered by data and the underlying dynamical system is changing with time. Conceptually, the attractor transition networks in the present study should be thought of not as representing the structure of an autonomous dynamical system, but rather the complex interaction between the environment and the brain dynamic landscape.
Our flexible, individualized construction of transition networks comes with a cost—there lacks a direct correspondence between different networks. An advantage of an HMM constructed from data concatenated across subjects (cf. Meer et al., 2020) is that the model comes with a fixed number of states across all subjects, with direct one-to-one correspondence. In contrast, the correspondence problem for the present method is data driven and needs to be solved separately. For example, attractor transition networks for different subjects (Figure 5A and B) contain different numbers of nodes (attractors) and edges (transitions). How nodes and edges in Figure 5A map to those in Figure 5B is not obvious. Even when comparing attractor transition networks with the same number of nodes, obtaining a correspondence is equivalent to solving an instance of the Graph Matching problem, which is a hard, fundamental challenge in graph processing (Umeyama, 1988; Zaslavskiy et al., 2009). In the present work, the correspondence between networks is made using techniques from optimal transport theory, specifically the use of Gromov–Wasserstein (GW) matchings between networks (Figure 3E and G) which can provide approximate graph matching solutions even for graphs with different numbers of nodes. GW matching does not require a temporal correspondence or any a priori knowledge of how the networks are constructed and can thus be used in a broad context. For example, for resting-state fMRI, there is no a priori way to map the transition network constructed from one recording session to that of another. GW matchings provide a solution to compare a subject’s resting transition networks in different sessions to examine the stability of the brain dynamic landscape, or to compare transition networks of healthy subjects to that of the psychiatric patients to examine the dynamic basis of the disorder. We therefore introduce this tool to the neuroimaging community with broader future applications in mind.
The present work also brings attention to the interpretation of dFC states. It is shown in Figure 4E and I and Figure S3D that dFC states are not equivalent to attractors in the brain dynamics: dFC is in part sensitive to the transitions (Figure 4I gray areas) and in part to the change in the control parameter without inducing any phase transition (Figure 4E white areas). It has been proposed that resting-state FC reflects cross-attractor transitions (Zhang et al., 2022). While dFC-based clustering can differentiate tasks very well (Gonzalez-Castillo et al., 2019), this differentiation may not be interpreted as the brain occupying distinct attractors during different tasks, but rather, these tasks involve different transitions and environment-driven correlations. Complementing existing dFC-based approaches, the attractor transition networks incorporate information of the underlying dynamical system and provide a channel to connect data-driven representations to dynamical systems concepts. Future work may further explore the relation between attractor transition networks and dFC by using single-frame techniques for FC quantification (Esfahlani et al., 2020; Faskowitz et al., 2020).
In application to the human fMRI data, we provide a series of analyses comparable to those of an earlier work (Saggar et al., 2018), where undirected networks were constructed using Mapper (Singh et al., 2007) from the same dataset (Gonzalez-Castillo et al., 2015). Saggar et al. (2018) observed that the core of the networks was dominated by cognitively demanding tasks, such as memory, math, and to a lesser extent, video tasks, while the peripheries were dominated by rest. In the same vein, we observe that the highest degree nodes are dominated by memory and math (Figure 5C), and that the level of dominance predicts behavioral performance (Figure 6A–C). Note that due to the compression step in the present construction (Figure 3B–C), tightly connected nodes in the neighborhood graph (Figure 3B) are contracted to a single node in the final attractor transition network (Figure 3C). It is reasonable to assume that tightly connected nodes that serve as the core for periphery nodes in the neighborhood graph would become a high-degree node in the compressed graph. Thus, the high-degree nodes in the present work may be thought of as a loose counterpart to the core of the Mapper-based construction (Saggar et al., 2018).
The periphery structures of interest in the present construction are the cycles that connect the high-degree nodes back to themselves. For the human fMRI data, long cycles are dominated by rest and the video task (Figure 5F), analogous to the Mapper-based result (Saggar et al., 2018) that periphery nodes are dominated by rest. Importantly, the present cycle-based analysis of the periphery structure allows us to examine recurrent dynamics of different duration (reflected as cycle length, Figure 5F) and identify the behavioral relevant time scales (Figure 6D–F). Interestingly, slower reaction time during tasks is associated with an excess of intermediate but not long cycles (Figure 6E). This suggests that intermediate cycles putatively represent the stray path that the brain took, resulting in slower reaction times. Interestingly, a greater number of cycles predicts higher accuracy, which, combined with the reaction time results, may reflect a speed-accuracy trade-off. It provides an example of process-based analysis of brain dynamics afforded by the present, dynamics-minded, construction. From a more technical viewpoint, the spectrum of cycle length (Figure 6D–F) is tightly connected to the multiscale nature of the attractor transition network. The Temporal Mapper naturally admits multiscale construction via the compression distance δ (Figure 3Aiv). That is, one can consider the attractor transition network as a sequence of networks at different scales δ (see Figure S9, for example) instead of a single graph. At a specific scale, any smaller recurrent processes, that is, smaller cycles formed by a few recurring attractors, will be contracted into a single node. Thus, the spectrum of cycle length indicates at which scales the graph can be further compressed. Further method development and empirical testing are required to better take advantage of the multiscale information in the Temporal Mapper. In addition, unfolding the dynamics in time as the average distance in the transition network tracks the transition between tasks (Figure 5G), comparable to Mapper-based results (Saggar et al., 2018). In the present work, the directed nature of the attractor transition networks introduces additional information—the direction of least action between states. Some nodes have a characteristic direction as a sink or a source (positive or negative deviation in Figure 5H), serving as entry and exit for the cognitively demanding tasks (Figure S8). Note that the sink-/source-ness of the associated brain activity patterns may not be absolute, as it is possible for the sink-/source-ness to depend on the specific design of the experiment (e.g., what types of tasks are included in the session and the time allocation for each task). Further studies are necessary to elucidate the interpretation of sink/sourceness in a more general context, which will require applying the Temporal Mapper to a wider range of datasets. Nevertheless, the present study demonstrates that the directedness introduced by the arrow time is task specific. Future work may further explore the cognitive correlates of different directed features of the graph, including the sink/sourceness of the nodes. Moreover, this directedness may be useful for designing neural stimulation protocols to more effectively perturb the brain into desirable states.
In conclusion, we propose a computational method for constructing attractor transition networks from simulated and empirical time series of brain dynamics. Complementing existing geometrical and statistical approaches to characterizing brain dynamics, the present work aims to provide a channel of correspondence between data-driven topological modeling and mechanistic modeling. Incorporating time in the construction of spatiotemporal neighborhoods, paths in the attractor transition networks encode the action of the underlying dynamical systems. The method is both validated using a biophysical network model of the brain and shown to reveal behavioral and cognitive relevant features in human fMRI data. The present work serves as a starting point for dynamical theory-driven topological analysis of brain dynamics. Future work will compare the present state and transition detection methods more extensively to existing community detection methods and further validate the method using consortium-sized data (Gordon et al., 2017; Saggar et al., 2022; Smith et al., 2013).
MATERIALS AND METHODS
Biophysical Network Model of the Human Brain
Parameter . | Interpretation . | Value . |
---|---|---|
τE | Decay time of NMDA receptor | 0.1 (s) |
τI | Decay time of GABA receptor | 0.01 (s) |
γE | Kinetic parameter of excitatory population | 0.641 |
γI | Kinetic parameter of inhibitory population | 1 |
aE | Parameter of HE | 310 (nC−1) |
bE | Parameter of HE | 125 (Hz) |
dE | Parameter of HE | 0.16 (s) |
aI | Parameter of HI | 615 (nC−1) |
bI | Parameter of HI | 177 (Hz) |
dI | Parameter of HI | 0.087 (s) |
rmax | Maximal firing rate | 500 (Hz) |
wEE | Excitatory-to-excitatory coupling | 2.8 (nA) |
wEI | Excitatory-to-inhibitory coupling | 1 (nA) |
wIE | Inhibitory-to-excitatory coupling | 2.8 (nA) |
wII | Inhibitory-to-inhibitory coupling | 0.05 (nA) |
II | External input to inhibitory population | 0.1 (nA) |
G | Global coupling | 1.1–5 (nA) |
Cij | Structural connectivity between brain regions | Human structural connectome (Civier et al., 2019; Van Essen et al., 2013) |
σ | Noise amplitude | 0.01 |
Parameter . | Interpretation . | Value . |
---|---|---|
τE | Decay time of NMDA receptor | 0.1 (s) |
τI | Decay time of GABA receptor | 0.01 (s) |
γE | Kinetic parameter of excitatory population | 0.641 |
γI | Kinetic parameter of inhibitory population | 1 |
aE | Parameter of HE | 310 (nC−1) |
bE | Parameter of HE | 125 (Hz) |
dE | Parameter of HE | 0.16 (s) |
aI | Parameter of HI | 615 (nC−1) |
bI | Parameter of HI | 177 (Hz) |
dI | Parameter of HI | 0.087 (s) |
rmax | Maximal firing rate | 500 (Hz) |
wEE | Excitatory-to-excitatory coupling | 2.8 (nA) |
wEI | Excitatory-to-inhibitory coupling | 1 (nA) |
wIE | Inhibitory-to-excitatory coupling | 2.8 (nA) |
wII | Inhibitory-to-inhibitory coupling | 0.05 (nA) |
II | External input to inhibitory population | 0.1 (nA) |
G | Global coupling | 1.1–5 (nA) |
Cij | Structural connectivity between brain regions | Human structural connectome (Civier et al., 2019; Van Essen et al., 2013) |
σ | Noise amplitude | 0.01 |
Note. Here we summarize the parameters used in the global model (Equation 4 and 5). Critical parameters are inherited from Wong and Wang (2006) to maintain biological plausibility.
Computation of the Attractor Repertoire and Ground-Truth Transition Network
Fixed points of the global model (Equations 4–6) are computed using fsolve in MATLAB for each parameter G (in 0.01 increments). A subset of the fixed points is further classified as attractors based on local stability analysis using the Jacobian matrix and simulated perturbation. These attractors are visualized in Figure 2D as a bifurcation diagram (for further computational details, see Zhang et al., 2022). Attractors that continuously map to each other under the variation of parameter G are considered qualitatively equivalent. Thus, in the present work, the term “attractor” is used in a broad sense, referring to the connected components of (narrow sense) attractors in the product of the state space and the parameter space. Computationally, single-linkage clustering is used to obtain these connected components. The set of all connected components or clusters constitutes the attractor repertoire, providing a compact description of the model dynamic landscape (Zhang et al., 2022).
The attractor repertoire further provides a skeleton to convert any simulated times series of the state variables (within the appropriate parameter range) into the corresponding symbolic dynamics. Here symbolic dynamics refers to a sequence of symbols, where each symbol represents an attractor, and the sequence represents the order in which the attractors are visited in time. Computationally, each time point in the simulated time series is assigned to an attractor using nearest neighbor classification in the product of the state space and the parameter space, given the precomputed attractor repertoire. The resulting symbolic dynamics is used to construct the ground-truth attractor transition network (Figure 2F). Nodes of the network are unique symbols appearing in the sequence. There is a link from node i to node j if the corresponding symbol i appears immediately before symbol j in the sequence. Each node is equipped with a probability measure proportional to the dwell time at the corresponding attractor (i.e., node size in Figure 2F).
Simulated Brain Activities and BOLD Signals
We simulate time series of brain activities and the derived BOLD signals as substrates for data-driven construction of attractor transition networks (see section Temporal Mapper—Construction of Transition Networks in Methods). The neural activity of each model brain region i is represented by the local excitatory population activity SE(i) (Equations 4–6), simulated using the Heun stochastic integration scheme with a 0.001 s time step.
Parameter . | Interpretation . | Value . |
---|---|---|
si | Vasodilatory signal | Variable |
fi | Blood inflow | Variable |
vi | Blood volume | Variable |
qi | Deoxyhemoglobin content | Variable |
κi | Rate of signal decay | 0.65 (s−1) |
γi | Rate of flow-dependent elimination | 0.41 (s−1) |
τi | Hemodynamic transit time | 0.98 (s) |
α | Grubb’s exponent | 0.32 |
ρ | Resting oxygen extraction fraction | 0.34 |
V0 | Resting blood volume fraction | 0.02 |
k1 | BOLD weight parameter | 7ρi |
k2 | BOLD weight parameter | 2 |
k3 | BOLD weight parameter | 2ρi − 0.2 |
Parameter . | Interpretation . | Value . |
---|---|---|
si | Vasodilatory signal | Variable |
fi | Blood inflow | Variable |
vi | Blood volume | Variable |
qi | Deoxyhemoglobin content | Variable |
κi | Rate of signal decay | 0.65 (s−1) |
γi | Rate of flow-dependent elimination | 0.41 (s−1) |
τi | Hemodynamic transit time | 0.98 (s) |
α | Grubb’s exponent | 0.32 |
ρ | Resting oxygen extraction fraction | 0.34 |
V0 | Resting blood volume fraction | 0.02 |
k1 | BOLD weight parameter | 7ρi |
k2 | BOLD weight parameter | 2 |
k3 | BOLD weight parameter | 2ρi − 0.2 |
Temporal Mapper—Construction of Transition Networks
The fundamental backbones of our transition networks are the k nearest-neighbor graphs (kNNG) that often appear in dimension reduction techniques (van der Maaten et al., 2009). Before introducing this construction, we first set up some preliminaries.
With this setup in place, we are ready to describe our construction. In our setting, data points that are close in time also have a tendency to be close in space, as the states change continuously (i.e., without sharp jumps) within an attractor. Keeping this tendency in mind, we carry out the following procedure:
Construct a standard kNNG
Remove edges of the form (xi, xi+1) (these will be added back in)
Remove all nonreciprocal connections, that is, only retain edges (xi, xj) if xi ∈ top(k, xj) and xj ∈ top(k, xi)
Add directed temporal links (xi, xi+1); existing undirected edges (xi, xj) are viewed as double edges (xi, xj), (xj, xi). This final digraph is denoted .
The initial pruning of temporal neighbors is carried out to help recover directionality in the data (see Figure S5). Other strategies may be possible, but the current method worked sufficiently well for our purposes. The intermediate step of removing nonreciprocal connections is a strategy for disambiguating between points that have differing local densities (Qin et al., 2011). The final addition of the temporal links injects the “arrow of time” back into the graphical representation. Note that this final step is not typical, but is possible in our setting because we work with time series data and know an explicit ordering of our data points. An important feature of the networks that we constructed was that they tended to be strongly connected as digraphs, meaning that it was possible to take a directed path between any two vertices. Our construction also accommodates multiple time series or time series with missing frames (frame censoring is common in fMRI time series to remove motion artifacts): the arrow of time is only removed and reintroduced between consecutive time points in the same time series or epoch in step 2 and 4 above. When the data constitute a single uninterrupted time series, the digraph will always have one connected component by virtue of the arrows of time that connect every pair of consecutive time points. When the data contain multiple disjoint time series, it is possible for the digraph to have multiple connected components for certain parameter k as the connectivity across time series depends solely on the existence of reciprocal spatial neighbors. If it is desirable for the digraph to have a single connected component, one can choose to (a) increase k until reciprocal kNN exists between any pair of time series, or (b) pre-align the time series to increase the spatial proximity between them.
Gromov–Wasserstein Distance Between Transition Networks
Temporal Mapper produces directed graphs where arrows display temporal structure. In general, comparing such graphs directly requires solving a correspondence problem as different graphs may have different numbers of nodes. To solve this correspondence problem, we use the Gromov–Wasserstein (GW) distance (Mémoli, 2007). While originally formulated for metric spaces, the GW formulation was shown to admit a bona fide distance between directed graphs with arbitrary edge weights in Chowdhury and Mémoli (2019) and has recently enjoyed significant attention from the machine learning community (Flamary et al., 2021; Peyré et al., 2016; Peyré & Cuturi, 2019; Solomon et al., 2016; Titouan et al., 2019; Xu et al., 2019). In the (di)graph setting, the GW distance allows one to compare the full structure of two (di)graphs without reducing to summary statistics such as degree distributions or centrality.
Because there are ∣V∣ × ∣W∣ individual entries making up ∣J∣, and the entries can all be computed independently, this problem is perfectly suited for GPU computations. Our GPU implementation of the TLB computes each coefficient Jij in a separate thread block so that all coefficients are computed in parallel. The final linear program can be solved easily using standard linear solvers, for example, using the network simplex algorithm (Bonneel et al., 2011). For applications in the present work, the GPU-infused version accelerates the original implementation (Chowdhury & Mémoli, 2019) by roughly 200 times.
Cycles of Transition Networks
Enumerating all cycles in a graph can be computationally expensive (Giscard et al., 2019). However, the transition networks that appear in our setting are small enough that we can enumerate all cycles and carry out further postprocessing in reasonable time. Given a transition network G with vertices indexed as {v1, v2, …, vd}, we proceed via the following heuristic approach. First we loop over all pairs of vertices (vi, vj) and use MATLAB’s native shortest path algorithm (i.e., a breadth-first search of complexity O(|V| + |E|)) to find the shortest paths from vi to vj and from vj to vi, respectively. These paths are then concatenated (avoiding trivial repetition at endpoints) to obtain a cycle. If a cycle has repeated nodes, that is, is not a simple cycle, then it is discarded. Finally, after the loop terminates, there may be multiple copies of the same cycle with different starting points. For each of these cases, we retain the copy starting at the smallest index and discard the others.
The Continuous Multitask Experiment
In this study, we utilized an fMRI dataset comprising 18 participants collected by Gonzalez-Castillo et al. (2015) by using a continuous multitask paradigm (CMP). We retrieved the data from the XNAT Central public repository (https://central.xnat.org; Project ID: FCStateClassif). Informed consent was obtained from all participants, and the local Institutional Review Board of the National Institute of Mental Health in Bethesda, MD, reviewed and approved the CMP data collection. We briefly describe the experiment structure and preprocessing below.
Participants were scanned continuously for 25 min and 24 s while performing four different cognitive tasks. Each task was presented for two separate 180-s blocks, and each task block was preceded by a 12-s instruction period. These four tasks were (a) Rest (R), where participants were instructed to fixate on a crosshair in the center of the screen and let their mind wander; (b) 2-back Working Memory (M), where participants were presented with a continuous sequence of individual geometric shapes and were instructed to press a button whenever the current shape was the same as the shape that appeared two shapes before; (c) Math/arithmetic (A), where participants were presented with simple arithmetic operations, involving three numbers between 1 and 10 and two operands (either addition or subtraction); and (d) Video (V), where participants watched a video of a fish tank from a single point of view with different types of fish swimming into an out of the frame, and were instructed to press a button when a red crosshair appeared on a clown fish and another when it appeared on any other type of fish. For arithmetic, the operations remained on the screen for 4 s, and successive trials were separated by a blank screen that appeared for 1 s, yielding a total of 36 operations per each 180-s block. For video, the targets appeared for 200 ms with a total of 16 targets during each of the 180-s blocks. The order of task blocks was randomized such that the same task did not appear in two consecutive blocks, and the same ordering of tasks was used for all participants. The randomized task order was R-M-V-A-M-R-A-V.
The fMRI data were acquired on a Siemens 7 Tesla MRI scanner equipped with a 32-channel head coil using a whole-brain echo planar imaging (EPI) sequence (repetition time [TR] = 1.5 s, echo time [TE] = 25 ms, and voxel size = 2 mm isotropic). A total of 1,017 volumes were acquired while participants performed the CMP.
Functional and anatomical MR images were preprocessed using the Configurable Pipeline for the Analysis of Connectomes (C-PAC version 0.3.4; https://fcp-indi.github.io/docs/user/index.html). We used the preprocessing utilized in a previous study (Saggar et al., 2018). Briefly, the fMRI data preprocessing steps included ANTS registration into MNI152 space, slice timing correction, motion correction, skull stripping, grand mean scaling, spatial smoothing (FWHM of 4 mm), and temporal band-pass filtering (0.009 Hz < f < 0.08 Hz). For each ROI, nuisance signal correction was performed by regressing out linear and quadratic trends, physiological noise (white matter and cerebrospinal fluid), motion-related noise (three translational and three rotational head-motion parameters) using the Volterra expansion (Friston et al., 1996) (i.e., six parameters, their temporal derivatives, and each of these values squared), and residual signal unrelated to neural activity extracted using the CompCor algorithm (Behzadi et al., 2007) (i.e., five principal components derived from noise regions in which the time series data were unlikely to be modulated by neural activity). The resulting data were brought to 3-mm MNI space, and the mean time series was extracted from 375 predefined regions of interest (ROIs) using the Shine et al. (2016) atlas. The atlas includes 333 cortical regions from the Gordon et al. (2016) atlas, 14 subcortical regions from the Harvard–Oxford subcortical atlas, and 28 cerebellar regions from the SUIT atlas (Diedrichsen et al., 2009). Individual ROIs with zero variance were excluded prior to computing attractor transition networks.
The behavioral data included both responses and reaction times for Working Memory, Math, and Video tasks. Participants were instructed to respond as quickly and accurately as possible with only one response per question. Behavior scores including the percent correct, percent missed, and response times for Working Memory (M), Math (A), and Video (V) tasks were computed for each participant.
SUPPORTING INFORMATION
Supporting information for this article is available at https://doi.org/10.1162/netn_a_00301. Custom MATLAB scripts used to generate simulated data and the implementation of Temporal Mapper is available at https://github.com/braindynamicslab/tmapper. The human fMRI data was originally collected by Gonzalez-Castillo et al. (2015) and is available for download from the XNAT Central public repository (https://central.xnat.org; Project ID: FCStateClassif).
AUTHOR CONTRIBUTIONS
Mengsen Zhang: Conceptualization; Formal analysis; Investigation; Methodology; Software; Visualization; Writing – original draft; Writing – review & editing. Samir Chowdhury: Conceptualization; Formal analysis; Methodology; Software; Validation; Visualization; Writing – original draft; Writing – review & editing. Manish Saggar: Conceptualization; Funding acquisition; Investigation; Methodology; Project administration; Supervision; Validation; Writing – review & editing.
FUNDING INFORMATION
Manish Saggar, National Institute of Mental Health (https://dx.doi.org/10.13039/100000025), Award ID: MH119735. Manish Saggar, Stanford Maternal and Child Health Research Institute (https://dx.doi.org/10.13039/100015521), Award ID: Faculty Scholar.
TECHNICAL TERMS
- Nonlinear dynamics:
The study of a type of dynamical systems, where dynamics of the whole system differs from that of independent parts.
- Attractors:
States and a set of states that attract nearby trajectories of the dynamical system. They are stable as the system returns to an attractor when slightly perturbed.
- Phase transitions:
Change in state of a dynamical system from the regime of one attractor to that of another either due to noise or underlying changes in the parameter of the system leading to destabilization of the current attractor.
- Multistability:
A feature of some nonlinear dynamical systems, where multiple attractors coexist in the state space.
- Metastability:
A type of dynamics in some nonlinear dynamical systems, where the system slows down (dwell) at specific preferred states and then escapes from it.
- Repellers:
States and a set of states that repel nearby trajectories of the dynamical system. They are unstable as the system moves further away from these states when slightly perturbed.
- Topological data analysis (TDA):
A field of applied mathematics that uses algebraic and computational techniques to represent data as topological objects and to compute global topological features of such objects.
- Mapper:
A TDA-based technique for generating graph representations that identify meaningful subgroups of high-dimensional data.
- Hysteresis:
A feature of multistable dynamical systems, where the system ends up in a different attractor than the one it started with as the control parameter changes continuously and returns to the same value.
- Control parameter:
A parameter of a dynamical system that is controlled or varied externally, which is not a state variable.
- Bifurcation:
A qualitative change in a dynamical system, where an attractor (repeller) is created, destroyed, or substantially modified by continuous changes in control parameter(s).
- Attractor transition network:
A directed graph represents the path of phase transitions in a dynamical system, whose nodes represent attractors visited and edges represent transitions occurred.
- Recurrence plot:
A N-by-N matrix for a time series of N samples, where element (i, j) describes the dissimilarity between the state of system at time i and at time j.
- K nearest neighbor (kNN) graph:
A graph built on point cloud data where nodes correspond to data points, and edges connect a node to its k (a positive integer) closest data points.
- Gromov–Wasserstein (GW) distance:
A measure of dissimilarity between two (possibly directed) graphs of possibly different sizes that optimizes over correspondences between the node sets of the two graphs.
- Reciprocal clustering of digraphs:
An agglomerative clustering technique that takes asymmetric edge weights into account when producing clusters.
REFERENCES
Author notes
Competing Interests: The authors have declared that no competing interests exist.
These authors contributed equally to this work.
Handling Editor: Christopher Honey