Systems neuroscience is facing an ever-growing mountain of data. Recent advances in protein engineering and microscopy have together led to a paradigm shift in neuroscience; using fluorescence, we can now image the activity of every neuron through the whole brain of behaving animals. Even in larger organisms, the number of neurons that we can record simultaneously is increasing exponentially with time. This increase in the dimensionality of the data is being met with an explosion of computational and mathematical methods, each using disparate terminology, distinct approaches, and diverse mathematical concepts. Here we collect, organize, and explain multiple data analysis techniques that have been, or could be, applied to whole-brain imaging, using larval zebrafish as an example model. We begin with methods such as linear regression that are designed to detect relations between two variables. Next, we progress through network science and applied topological methods, which focus on the patterns of relations among many variables. Finally, we highlight the potential of generative models that could provide testable hypotheses on wiring rules and network progression through time, or disease progression. While we use examples of imaging from larval zebrafish, these approaches are suitable for any population-scale neural network modeling, and indeed, to applications beyond systems neuroscience. Computational approaches from network science and applied topology are not limited to larval zebrafish, or even to systems neuroscience, and we therefore conclude with a discussion of how such methods can be applied to diverse problems across the biological sciences.

How can we make sense of the brain? Researchers can now record the activity of thousands of individual neurons across the whole brain (Figure 1A) thanks to methodological improvements in multielectrode arrays (Branner & Normann, 2000; Paulk et al., 2022; Steinmetz et al., 2019), and calcium indicators combined with fast microscopy (Aimon et al., 2019; T.-W. Chen et al., 2013; Hastings et al., 1969; Kim et al., 2017; Miyawaki et al., 1997; Nakai et al., 2001; Nguyen et al., 2016; Stringer et al., 2019; Tsien et al., 1982; Vanwalleghem et al., 2018). With these new tools in hand, the number of recorded neurons has doubled every seven years since 1960 and, as a consequence, neuroscience is relying increasingly on “big data” approaches (I. H. Stevenson & Kording, 2011). While the methodological advances that allow researchers to fill the space between individual neurons and the whole brain are exhilarating, the mountains of data behind them present new conceptual challenges to understand them fully.

Figure 1.

Rapid progress of systems neuroscience and the encoding of stimuli. (A) Across animal models of increasing complexity, we are now able to record the activity of many neurons. For example, we can record all of the neurons of C. elegans, about 2,000 neurons of D. melanogaster (Tainton-Heap et al., 2021), and about 80% of neurons in the larval zebrafish (D. rerio; X. Chen et al., 2018). High-density electrophysiological probes allow the recording of hundreds of units per session in mice, including deep structures within the brain (Steinmetz et al., 2019); in fact, researchers have recorded up to 10,000 neurons in a 0.3 mm3 portion of the brain using two-photon imaging (Stringer et al., 2019). (B). Stimulus encoding can occur at the single neuron level, where, for example, it could encode the direction of a stimulus. Stimulus encoding can also occur at the population level, where multiple units together can provide a better representation of the stimulus, including its direction and color. (C) Example of populations of neurons responding to the direction of water flow along the fish’s tail (Vanwalleghem et al., 2020). These neuronal populations show a tonic activity for the duration (10 s) of a water flow stimulus from the tail to the head (green arrow), or from the head to the tail (magenta arrow). The neurons were clustered in three categories: bidirectional response (blue), head to tail (pink), or tail to head (gold). Here we represent their spatial coordinates in the zebrafish brain (left). Telen = telencephalon, OT = optic tectum, TS = torus semicircularis (small dotted lines), ON = octavolateralis nucleus. (D) It is possible to predict neuronal activity from behavior, which in this example is the facial movement of a mouse. (Top) Motion energy was computed from consecutive frames (t and t + 1). (Middle) The principal components could then be used to predict the neuronal activity of 1,000 neurons. (Bottom) The predicted activity was remarkably similar (about 30%) to the real neuronal activity (sorted with an embedding for visualization). Adapted from Stringer et al. (2019).

Figure 1.

Rapid progress of systems neuroscience and the encoding of stimuli. (A) Across animal models of increasing complexity, we are now able to record the activity of many neurons. For example, we can record all of the neurons of C. elegans, about 2,000 neurons of D. melanogaster (Tainton-Heap et al., 2021), and about 80% of neurons in the larval zebrafish (D. rerio; X. Chen et al., 2018). High-density electrophysiological probes allow the recording of hundreds of units per session in mice, including deep structures within the brain (Steinmetz et al., 2019); in fact, researchers have recorded up to 10,000 neurons in a 0.3 mm3 portion of the brain using two-photon imaging (Stringer et al., 2019). (B). Stimulus encoding can occur at the single neuron level, where, for example, it could encode the direction of a stimulus. Stimulus encoding can also occur at the population level, where multiple units together can provide a better representation of the stimulus, including its direction and color. (C) Example of populations of neurons responding to the direction of water flow along the fish’s tail (Vanwalleghem et al., 2020). These neuronal populations show a tonic activity for the duration (10 s) of a water flow stimulus from the tail to the head (green arrow), or from the head to the tail (magenta arrow). The neurons were clustered in three categories: bidirectional response (blue), head to tail (pink), or tail to head (gold). Here we represent their spatial coordinates in the zebrafish brain (left). Telen = telencephalon, OT = optic tectum, TS = torus semicircularis (small dotted lines), ON = octavolateralis nucleus. (D) It is possible to predict neuronal activity from behavior, which in this example is the facial movement of a mouse. (Top) Motion energy was computed from consecutive frames (t and t + 1). (Middle) The principal components could then be used to predict the neuronal activity of 1,000 neurons. (Bottom) The predicted activity was remarkably similar (about 30%) to the real neuronal activity (sorted with an embedding for visualization). Adapted from Stringer et al. (2019).

Close modal

Systems neuroscience is less interested in the activity, either spontaneous or elicited, of each individual neuron, but more in how ensembles, or populations, of neurons form and interact (Figure 1B). Population coding complements the single neuron encoding approach (Barlow, 1972; Theunissen & Miller, 1995), with a population-level encoding of information by ensembles of neurons (Knight, 1972). This population level of encoding can combine multiple features of a stimulus and ensure the robustness of a behavior despite individual neurons’ stochasticity and noise (Quian Quiroga & Panzeri, 2009). Of course, population-level encoding can only be described and analyzed if one can record from multiple neurons simultaneously.

A number of approaches have been developed or adapted for the analysis of population-level data. These approaches include dimensionality reduction (Mu et al., 2019; A. W. Thompson et al., 2016), linear regression (X. Chen et al., 2018; Vladimirov et al., 2018), clustering (Mölter et al., 2018; Vanwalleghem et al., 2020), nonlinear analyses (Ponce-Alvarez et al., 2018), and topological approaches (Gracia-Tabuenca et al., 2020; A. E. Sizemore et al., 2019), to name a few. For many of these methods, the dimensionality and size of population-level datasets can be a limiting factor and require either the use of distributed computing (Freeman et al., 2014), dimensionality reduction (Betzel, 2020), or new mathematical tools (Haesemeyer et al., 2019; Tajima et al., 2015; Triplett et al., 2020; Watanakeesuntorn et al., 2020). Another challenge is the fabled needle in a haystack: identifying a potential sparse population that would drive the animal’s response (Ince et al., 2013; Tajima et al., 2017). Ultimately, all of these tools aim to understand how the brain encodes stimuli and transforms the sensory input into meaningful information and behavior.

Neuronal ensembles encoding sensory stimuli have been observed in zebrafish for all sensory modalities from visual, to auditory and vestibular (Avitan et al., 2020; X. Chen et al., 2018; Favre-Bulle et al., 2020; Favre-Bulle et al., 2018; Friedrich et al., 2004; Migault et al., 2018; Poulsen et al., 2021; A. W. Thompson et al., 2016; Vanwalleghem et al., 2017; Vanwalleghem et al., 2020). For example, the lateral line is a specialized apparatus that senses water flow in fish and amphibians. We showed the existence of neuronal representations of the direction and speed of the water flow (Figure 1C), as well as of the simulated distance traveled by the animal (Vanwalleghem et al., 2020). For visual processing, X. Chen et al. (2018) showed how the transformation of visual stimuli into motor commands was distributed across the brain, before converging onto small clusters of neurons in the hindbrain that output the motor commands. A systems-wide approach was needed to observe the distributed nature of neuronal representations in the above studies and has shown a similar distribution across the brain in mice using Neuropixels probes (Steinmetz et al., 2019) and across the visual cortex using two-photon imaging (Figure 1D; Stringer et al., 2019).

The key limitation of most whole-brain work is its descriptive nature. Studies have linked neuronal responses to stimuli, but few offer testable models or hypotheses on how the sensory processing can drive behavior. Testing the causal link between neuronal responses and behavior is particularly complex when the responses are spatially distributed, as described above. Promising approaches for surmounting this challenge include light-shaping and optogenetics, or ablation, which could allow neuroscientists to interrogate distributed neuronal circuits and identify the necessary parts underlying a specific behavior (Dal Maschio et al., 2017; Dunn et al., 2016; Fernandes et al., 2021; Marshel et al., 2019; Naumann et al., 2016; Vladimirov et al., 2018). For example, Helmbrecht et al. (2018) showed that optogenetic activation of specific regions within the optic tectum could trigger different behavioral programs: an approach program or an escape attempt. The investigators further identified multiple classes of projection neurons from the optic tectum to the hindbrain. These projections could convey the relevant information to the motor centers, including the position of the stimulus. Such a combination of structural and functional data provides a more comprehensive picture of, as well as testable hypotheses regarding, how information flows between brain regions and how that information could be transformed to drive specific behaviors.

Our goal is for this review to be accessible to biologists with little or no mathematical training, since understanding the basis of the most common approaches will empower neuroscientists to choose the most suitable method to answer their specific research question. As most of the methods we discuss involve relations between units, we will begin with approaches for quantifying simple relations between two variables, and from there build up to approaches involving multiple variables such as network science. Finally, we will examine the potential of generative models, and discuss avenues for future research, and how these tools can be used in other contexts. We direct the readers who want to deepen their understanding of a specific topic to these excellent reviews on linear models for encoding (Marinescu et al., 2018; Paninski et al., 2007) and graph theory (Bassett & Sporns, 2017; Sporns, 2018).

Systems neuroscience offers the perspective that no one neuron functions alone, visual stimuli influence future motor sequences, a triggered memory may evoke an emotional reaction, and sudden sounds shift attention. All of these examples can be simplified into one unit, for example a neuron or population (see Figure 2A), that influences, connects to, or relates to another unit. Determining whether two units are related may seem to be simple at first glance. For example, in which periods of time (Figure 2, x-axis) is the activity time series of one unit tracking the activity of the other unit? Despite the simple question statement, the path to an answer is not immediately obvious, as there are many types of computational methods available from which we must choose. We will highlight a few classical examples, including linear regression and causal inferences, but invite interested readers to peruse Cliff et al. (2022) for a deeper comparison.

Figure 2.

Calculating relationships between two units. (A) In systems neuroscience, a single unit may correspond to a neuron or a population of neurons. Two units may potentially interact in a variety of ways. For example, one can either activate or inhibit the other; the strength of these interactions can also vary, and feedback between units can also exist. As an example, units x (blue) and y (orange) have activities that change over time. At time i, unit x has activity xi and unit y has activity yi. (B) Multiple methods exist for quantifying relations between two variables. (Left) Pairwise similarity measures often see the leftmost portion of the activity traces as having high similarity and the middle portion as having low similarity; these measures may vary on their interpretation of the rightmost portion. (Middle) Linear regression calculates a line of best fit (darker gold) given the points (xi, yi) for each time i. This line of best fit minimizes the error between the predicted values of yi, using xi, and the true values yi (lighter gold). (Right) Causal inference methods ask whether information about the activity of unit x between time ik and the current time i improves the prediction accuracy of the future activity of unit y, above and beyond the prediction accuracy obtained by using the previous activity of unit y alone. (C) Applications of each method to larval zebrafish data. (Left) The pairwise similarity (correlation) between spinal neurons increases dramatically between 18 and 20 hours post fertilization (hpf), as can be seen by the blue and red activity traces moving in synchrony. Adapted from Warp et al. (2012). (Middle) Linear regression was used to quantify the responses to auditory stimuli; the coefficient is mapped to the color of each neuron in the brain, while the coefficient of determination R2 is mapped to the size of the sphere. Adapted from Poulsen et al. (2021). (Right) The Granger causality approach was used to identify the flow of information between auditory brain regions in Vanwalleghem et al. (2017). Th = thalamus, TS = torus semicircularis, ON = octavolateralis nucleus, Hb = hindbrain.

Figure 2.

Calculating relationships between two units. (A) In systems neuroscience, a single unit may correspond to a neuron or a population of neurons. Two units may potentially interact in a variety of ways. For example, one can either activate or inhibit the other; the strength of these interactions can also vary, and feedback between units can also exist. As an example, units x (blue) and y (orange) have activities that change over time. At time i, unit x has activity xi and unit y has activity yi. (B) Multiple methods exist for quantifying relations between two variables. (Left) Pairwise similarity measures often see the leftmost portion of the activity traces as having high similarity and the middle portion as having low similarity; these measures may vary on their interpretation of the rightmost portion. (Middle) Linear regression calculates a line of best fit (darker gold) given the points (xi, yi) for each time i. This line of best fit minimizes the error between the predicted values of yi, using xi, and the true values yi (lighter gold). (Right) Causal inference methods ask whether information about the activity of unit x between time ik and the current time i improves the prediction accuracy of the future activity of unit y, above and beyond the prediction accuracy obtained by using the previous activity of unit y alone. (C) Applications of each method to larval zebrafish data. (Left) The pairwise similarity (correlation) between spinal neurons increases dramatically between 18 and 20 hours post fertilization (hpf), as can be seen by the blue and red activity traces moving in synchrony. Adapted from Warp et al. (2012). (Middle) Linear regression was used to quantify the responses to auditory stimuli; the coefficient is mapped to the color of each neuron in the brain, while the coefficient of determination R2 is mapped to the size of the sphere. Adapted from Poulsen et al. (2021). (Right) The Granger causality approach was used to identify the flow of information between auditory brain regions in Vanwalleghem et al. (2017). Th = thalamus, TS = torus semicircularis, ON = octavolateralis nucleus, Hb = hindbrain.

Close modal

Is the Activity of Two Units Pairwise Similar?

The precise definition of pairwise similar has resulted in a multitude of mathematical descriptions that we will not explore here (Bobadilla-Suarez et al., 2020; Miri et al., 2011). At a basic level, the activity of two units being similar might mean that changes in the activity of unit x coincide with changes in the activity of unit y (Figure 2B). For example, the correlation between spinal neurons drastically increases during the development of larval zebrafish (Figure 2C, left; Warp et al., 2012). The central pattern generator is dependent on this correlated, and contralaterally anticorrelated, activity to generate the rhythmic movements mediating swimming in zebrafish. Importantly there are myriad additional approaches to the problem of evaluating a relationship between two units (Cutts & Eglen, 2014). While no one method is error-free, supporting conclusions with multiple types of pairwise similarity measures from different domains can reduce the effects of bias on the overall outcome.

How Is the Activity of Two Units Related?

Given the activity of two neurons, we might want to understand how they are related. Linear regression can be used to answer this question, by expressing the activity of the second neuron as a linear function of the activity of the first neuron (Figure 2B, middle). When additional variables are present that might affect the results, such as neuron type or location, we can also incorporate these variables into our model to understand or mitigate the influence of such variables on the relationship in question (Smith, 2018). Linear regression is a common but powerful tool that sets the stage for more general approaches such as general linear models and machine learning. In systems neuroscience specifically, these techniques have been used to assess learning (Luchiari & Chacon, 2013; Tompson et al., 2020) and to determine neurons associated with a behavior (Marques et al., 2020; Miri et al., 2011). Linear regression has also been used to assess the response to auditory stimuli in larval zebrafish (Figure 2C, middle). Poulsen et al. (2021) showed that the auditory responses were more widespread than previously observed, but also confirmed previous observations of the primary auditory regions in larval zebrafish (Privat et al., 2019; Vanwalleghem et al., 2017).

Causal Inference

Knowledge of an association does not necessarily help us with designing intervention strategies, mapping out novel circuits, or determining mechanisms (Nienborg & Cumming, 2010). In order to reach these goals, we need to start with an understanding of causal relationships between neuronal units. For this review, we will say a cause of an effect must both precede the effect and provide predictive information about the effect (Cekic et al., 2018; Granger, 1969; Wiener, 1956). This notion of a cause is commonly referred to as Granger causality or G-causality. If unit x activity causes a change in unit y activity, then given our perspective on causality, we would expect to see that past and present activity of unit x can predict the activity of unit y, or that unit x activity Granger-causes unit y activity. These ideas support the way many expect the brain to function, and so have been used frequently in neuroscience. For example, G-causality has been used to understand information flow in auditory processing, which confirmed previous results of neuroanatomical studies and aligned with evidence in the mammalian literature (Figure 2C, right; Vanwalleghem et al., 2017) and prey capture (Oldfield et al., 2020) in larval zebrafish. G-causality is one among many causal inference approaches; for example, a method was specifically designed to infer causality using optogenetics by compensating for the inherent confounds (Lepperød et al., 2018). Other methods, such as empirical dynamic modeling, are more appropriate for nonlinear dynamical systems and have been applied to neuronal data (Watanakeesuntorn et al., 2020).

In summary, there are many ways to test for the presence of a relationship between two interacting units, and while the methods we discussed have been useful for systems neuroscience, it is important to choose the approach based on the data, the system, and the specific question at hand (Torres et al., 2020).

Instead of focusing on only one pair of variables, we can expand our scope to include many interacting units such as thousands of neurons or hundreds of neuronal populations (Figure 3A). Approaches under the umbrella of complexity, or network, science mathematically aim to encode the data as units and parts (interactions, relations) of a quantitative object, such as a network, then use computational methods to evaluate the representation and make predictions about how that structure influences biological function. Below we draw mainly from the two subfields of network neuroscience and topological data analysis, but we invite the interested reader to peruse Battiston et al. (2020), Spivak (2009), and Torres et al. (2020) for deeper discussions.

Figure 3.

Systems neuroscience uses multiple methods to encode, represent, and analyze data. (A) Data consists of many units with recorded activity over time. (B) There are many options for encoding data into a mathematical representation. (C) Examples of representations. (Left) A network is composed of nodes and edges that connect exactly two nodes. (Middle) A simplicial complex is constructed from nodes and simplices that connect any number of nodes, and any subset of connected nodes. (Right) A hypergraph is built from nodes and hyperedges that connect any number of nodes. (D) Based on the type of representation used (network, simplicial complex, hypergraph), different analysis approaches become available. For example, representing data as a network allows one to detect many types of structure, and more recently very specific patterns such as core-periphery organization. Representing data as a simplicial complex permits the use of homology, which perceives the circular nature of the complex. Finally, a hypergraph representation enables a unique perspective on the community structure of the system. While one could use any of these representations to calculate, for example, community structure, we stress that each representation will provide a unique perspective on community structure, and furthermore that some representations are more amenable to particular analyses, for example, finding loops, than others (Torres et al., 2020). (E) Application of network theory to larval zebrafish data. WT and fmr1-/- larval zebrafish were presented with increasing intensity of sound, from −21 dB to −9 dB, and their neuronal activity was recorded. Circle plots of the edges between brain region nodes in WT and fmr1-/- zebrafish; an edge was placed if the correlation between their response to auditory stimuli is above 0.85. Node color indicates brain regions, in order: telencephalon, red; torus semicircularis, dark magenta; cerebellum, dark green; thalamus, orange; hindbrain without the Cb and ON, gray; octavolateralis nucleus, magenta; pretectum, light blue; optic tectum, blue; habenulae, yellow; tegmentum, light green. Adapted from Constantin et al. (2020). (F) Application of topological data analysis to the structural connectome. Example conserved cavity connecting visual processing regions shown in the brain (left) and abstracted for ease of interpretation (right). For visual simplicity, the cavity depiction in the brain does not include shading for 2-simplices, only edges. Adapted from A. E. Sizemore et al. (2018). (G) A hypergraph created from recordings of the on/off activity of brain regions during four tasks reveals that the green node A is only active in groups of three regions, whereas the blue node D is only active in groups of two regions. As such the green, magenta, and red nodes are all linked by a hyperedge. Adapted from Torres et al. (2020).

Figure 3.

Systems neuroscience uses multiple methods to encode, represent, and analyze data. (A) Data consists of many units with recorded activity over time. (B) There are many options for encoding data into a mathematical representation. (C) Examples of representations. (Left) A network is composed of nodes and edges that connect exactly two nodes. (Middle) A simplicial complex is constructed from nodes and simplices that connect any number of nodes, and any subset of connected nodes. (Right) A hypergraph is built from nodes and hyperedges that connect any number of nodes. (D) Based on the type of representation used (network, simplicial complex, hypergraph), different analysis approaches become available. For example, representing data as a network allows one to detect many types of structure, and more recently very specific patterns such as core-periphery organization. Representing data as a simplicial complex permits the use of homology, which perceives the circular nature of the complex. Finally, a hypergraph representation enables a unique perspective on the community structure of the system. While one could use any of these representations to calculate, for example, community structure, we stress that each representation will provide a unique perspective on community structure, and furthermore that some representations are more amenable to particular analyses, for example, finding loops, than others (Torres et al., 2020). (E) Application of network theory to larval zebrafish data. WT and fmr1-/- larval zebrafish were presented with increasing intensity of sound, from −21 dB to −9 dB, and their neuronal activity was recorded. Circle plots of the edges between brain region nodes in WT and fmr1-/- zebrafish; an edge was placed if the correlation between their response to auditory stimuli is above 0.85. Node color indicates brain regions, in order: telencephalon, red; torus semicircularis, dark magenta; cerebellum, dark green; thalamus, orange; hindbrain without the Cb and ON, gray; octavolateralis nucleus, magenta; pretectum, light blue; optic tectum, blue; habenulae, yellow; tegmentum, light green. Adapted from Constantin et al. (2020). (F) Application of topological data analysis to the structural connectome. Example conserved cavity connecting visual processing regions shown in the brain (left) and abstracted for ease of interpretation (right). For visual simplicity, the cavity depiction in the brain does not include shading for 2-simplices, only edges. Adapted from A. E. Sizemore et al. (2018). (G) A hypergraph created from recordings of the on/off activity of brain regions during four tasks reveals that the green node A is only active in groups of three regions, whereas the blue node D is only active in groups of two regions. As such the green, magenta, and red nodes are all linked by a hyperedge. Adapted from Torres et al. (2020).

Close modal

Encoding Data

The first step is to decide what the units, or nodes, will be, either individual neuronal units, or nodes representing multiple neuronal units, such as those that respond to the same stimulus (Figure 3A; Crossley et al., 2013). The choice of encoding method, which takes data and creates from it a quantitative object (Figure 3B), is crucial to the success of the entire analysis. If chosen incorrectly, it can cause unexpected results, bias the computations, and prevent the analysis from answering the initial question (Torres et al., 2020). Finally, we must decide what will constitute a relation between units. As discussed above, one could use pairwise similarity, or causal inference to identify relations between any pair of nodes (Fallani et al., 2015; Watanakeesuntorn et al., 2020). Using these ideas, one can construct a graphical model in which a relation between two units indicates a pairwise dependence. Such models are common in neuroscience (Belilovsky et al., 2015; Chang et al., 2019). Additionally, techniques from machine learning can provide a whole-brain approach to creating networks (Murugesan et al., 2020). Finally, one can use probabilistic graphical models to represent neuronal ensembles and identify neurons that can activate the whole ensembles (Carrillo-Reid et al., 2021).

Many of the above discussed methods use data only from the two units in question, but when encoding data involving many units, we can use methods that incorporate all of these data. For example, Bayesian networks (Bielza & Larrañaga, 2014; Javidian et al., 2020) provide a representation of the joint probability distributions across all the units within the system, so that a relation indicates dependence of the target on the source. Finally, a method called Mapper creates a representation from high-dimensional data by forming nodes from (overlapping) clusters of data points and relations between nodes with corresponding clusters that share data points (Cámara, 2017; Singh et al., 2007). Mapper’s representation not only offers a new way to visualize high-dimensional data, but additionally creates a computational object (network) upon which one can perform downstream analyses such as assessing community structure of a Mapper representation created from fMRI data (Saggar et al., 2018).

Importantly, the type of object constructed from the data-encoding process can be constrained by the method of encoding. Let us consider three mathematical frameworks with which we could represent our data (Figure 3C). The first is a network, which has units (nodes) and exclusively pairwise relations between nodes. The second and third, simplicial complexes and hypergraphs, respectively, allow for polyadic relations among nodes, meaning that any number of nodes can be related. Taken together, in choosing the encoding strategy it is crucial to consider (a) how to define nodes, (b) what a relation between nodes should capture, and (c) the frameworks that can be created with the approach. It is then important to ensure that all of these fit within the larger research question and biological system.

Network Science

By representing data as a network composed of nodes interconnected by edges, the investigator can answer new questions outside the reach of analytical techniques that focus on pairwise interactions (Bassett et al., 2018). Specifically, one can ask questions about how the full pattern of relations—the network—might relate to the system’s dynamics, the associated processes of perception, action, and cognition, and the resultant behavior (Fornito et al., 2016). Across many networks, the values of related measures are correlated; notably, however, the converse is not necessarily true (da Fontoura Costa et al., 2007). The values of mathematically unrelated measures can still be correlated across a network ensemble if they are being simultaneously driven by a single mechanism (Onnela et al., 2012; A. Sizemore et al., 2016). Relatedly, some statistical measures might vary across an ensemble in a way that tracks variation in health, severity of disease, type of computation, and accuracy, complexity, or form of behavior.

One dimension along which network measures range is the dimension of scale, from local, as typically thought to be reflected in the clustering coefficient, to global, as typically thought to be reflected in the shortest path length (Bullmore & Sporns, 2009). Networks that have both high clustering (similar to a lattice) and low shortest path length (similar to a random network) are said to have small-world organization (Bassett & Bullmore, 2006, 2017; Muldoon et al., 2016). Evidence suggests that the network of neurons in zebrafish larvae has a small-world organization (Stobb et al., 2012), suggesting its capacity for both local information processing and global information transmission. This organization is also seen in C. elegans and other species across diverse evolutionary lineages (Kaiser & Varier, 2011). However, it may be that this small-world organization only holds at certain scales (synapses, circuits, brain areas, etc.), or depends on the empirical approach taken to gather and analyze the data (Hilgetag & Goulas, 2016; Muller et al., 2014). For example, scale-free networks, where the degree distribution follows a power law, have also been observed in the optic tectum of larval zebrafish (Avitan et al., 2017), and biological networks as a whole have been found to exhibit strong to strongest evidence of scale-free structure in 12% of the networks studied (Broido & Clauset, 2019). It will be important for future generative models to incorporate additional factors beyond small-worldness, such as the importance of long-term connections to the functional diversity of distant brain areas, and not only topological considerations (Betzel & Bassett, 2018).

In the middle of the continuum from local to global structure in networks is the mesoscale (Figure 3D); for example, a network is said to have core-periphery structure if it contains (a) a core, which is a set of densely interconnected nodes, and (b) a periphery, which is a set of nodes that connect to the core, but do not tend to connect to other nodes in the periphery (Rombach et al., 2014). It is important to note that the local and global measures discussed above are independent of the structure. Notably, mesoscale structure has been shown to vary across cognitive states (Pedersen et al., 2018) and stages of learning (Bassett & Mattar, 2017), and to be altered by genetic risk for psychiatric disease (Dimitriadis et al., 2021). These mesoscale structures have been observed in Drosophila (Shih et al., 2015; Shih et al., 2020); functional connectomes display hierarchical modularity that reconfigures appreciably in response to a stimulus in larval zebrafish (Betzel, 2020). Constantin et al. (2020) showed that the zebrafish model of fragile X syndrome had increased sensitivity to sound, especially apparent in the higher connectivity across brain regions at lower sound intensity (Figure 3E, −15 dB). Such an increased connectivity is consistent with one of the theories developed to explain the basis of the sensory phenotype in fragile X syndrome (Testa-Silva et al., 2012). For a deeper dive in the use of networks with calcium imaging data, we invite readers to consider Nelson and Bonner (2021).

Topological Data Analysis

Methods from topological data analysis (TDA) aim to provide a quantitative and qualitative description of how the pieces of the system combine to form the larger whole (Blevins & Bassett, 2020; Geniesse et al., 2019; Ghrist, 2007; Munch, 2017). This topological perspective deviates from many familiar frameworks, such as principal component analysis, by requiring no a priori assumption about the space of the data (for example, that the data are linear). Similar to network science, TDA can represent systems with a simplicial complex, which contains nodes (units) and simplices (parts or interaction representations, see Figure 3C). While the edge of a graph can connect only two nodes, a simplex can connect any number of nodes. For example, three nodes can be all together connected by a simplex (illustrated as a filled triangle in Figure 3C); specifically, a k-simplex connects k + 1 nodes in a simplicial complex. In Figure 3C, we see that the simplicial complex has simplices drawn as links (1-simplices), shaded triangles (2-simplices), and shaded tetrahedra (4-simplices), and thus can represent many more types of multi-node interactions than the network. Encoding such polyadic relations between nodes provides TDA with distinct advantages, such as the ability to represent many-unit neuronal interactions and the means to detect system features not perceived in other domains. If we examine three neurons responding to different stimuli, the simplicial complex lets us represent the difference between a system in which every pair of neurons co-fires in response to different stimuli, and a system in which all three neurons fire in response to the same stimulus. Though we here have presented only a brief introduction to these TDA ideas, we direct the interested reader to Ahissar and Kleinfeld (2003), Blevins and Bassett (2020), Lisman and Grace (2005), Ma’ayan et al. (2008), Munch (2017), A. E. Sizemore et al. (2019), and Tallon-Baudry et al. (2001).

Loops within a neuronal system can support memories (Lisman & Grace, 2005; Tallon-Baudry et al., 2001), permit computations (Ahissar & Kleinfeld, 2003), and affect dynamics (Ma’ayan et al., 2008). We can detect such looped motifs by using homology to detect loops that surround topological cavities or voids within the simplicial complex representation of the system (for example, the large looped nature of the simplicial complex in Figure 3D). Specifically, loops that surround a cavity would intuitively allow for two routes of information to travel from one side of the loop to another, or could allow a signal to cyclically continue around the loop and thus repeat periodically. Furthermore, homology, and its weighted counterpart, persistent homology, can detect cavities, and thus loops, of any dimension within the system. For example, Reimann et al. (2017) identify high-dimensional loops within the correlated activity of neurons, and find that in general such topological complexity depends upon neuron morphology. Additionally, A. E. Sizemore et al. (2018) found loops connecting cortical to subcortical regions in the structural connectome to be a consistent feature across healthy adults. Shown in Figure 3F, one of these conserved loops connects higher order visual processing regions and mimics portions of the dorsal and ventral visual streams. The conserved nature of these loops found across individuals, but not in null models, suggests their importance in computation (A. E. Sizemore et al., 2018). Additionally, these homological ideas have improved our understanding of neuronal dynamics (Bardin et al., 2019; Curto, 2016; Curto et al., 2012), stimulus encoding (Basso et al., 2016; Giusti et al., 2015; Rubin et al., 2019), and the time-evolving brain (Yoo et al., 2016).

Although we introduced Mapper as a method for encoding data, it has also been used as a stand-alone tool to understand neural systems as it allows one to represent large datasets in a reduced form. Simply having the ability to visualize high-dimensional data using the Mapper algorithm has led to the discovery of a new breast cancer subgroup (Nicolau et al., 2011) and has been used to predict drug effects in traumatic brain injury (Nielson et al., 2015). Additionally, Saggar et al. (2018) paired Mapper with functional imaging data to investigate features of an individual’s brain state space. Their work led to the creation of DyNeuSR (Geniesse et al., 2019), a Python implementation of Mapper designed specifically for neuroimaging. With these implementations of Mapper, researchers can get a graphical representation of brain states, for example, observe how individuals traverse across the graph (brain states), and from this sequence better understand state transitions between different cognitive tasks (Geniesse et al., 2022).

Alternative Representations

Not all neural systems are most faithfully represented by networks or simplicial complexes. Many other representations exist, including but not limited to hypergraphs, higher order networks, multilayer networks, and sheaves, which we will all briefly describe below. Similar to simplicial complexes, hypergraphs can encode polyadic relationships among nodes (in Figure 3C, comparing simplices to hyperedges). However, whereas a 2-simplex (a shaded triangle) implies that all pairs of nodes are pairwise related, a hyperedge connecting three nodes does not imply that any pair of the three nodes is pairwise related. This feature allows the hypergraph to represent a wide variety of systems, such as functional (Davison et al., 2015; Davison et al., 2016) and structural neural systems. For example, after recording neural responses across different tasks, we could create a hypergraph in which brain regions are connected by a hyperedge if they are coactive in a task (Figure 3G). For example, the three regions A, B, C are all coactive in a task (Figure 3G, task 1), but regions A and B are never coactive in any other task without C. Detecting many large hyperedges (hyperedges that connect many nodes) indicates that often many populations act in unison in the tasks. Furthermore, if we find many large hyperedges but few small hyperedges, then we could conclude that generally these populations function only alongside others, and rarely by themselves or in small groups. Such a phenomenon can be quantified by the hypergraph-specific measurement called the fill coefficient (Torres et al., 2020).

Higher order networks (Xu et al., 2016) are designed to handle data from paths, where the activity of a node could depend on multiple previous events, such as from a sequence of visited brain states where the current brain state does not depend on only the one immediately preceding it. For example, in Figure 3G one could imagine a situation in which the response of region B to task 4 would depend on the responses of region D not only in task 2, but also in task 1. Such a situation would show a second-order dependency of B as it depends on the state of D two steps back. Multilayer networks represent systems that might (a) have different types of edges such as correlations in different frequency bands (Buldú & Porter, 2018), or different data types entirely; (b) evolve with time (W. H. Thompson et al., 2017); or (c) include multiple subjects (Muldoon & Bassett, 2016). For example, one could imagine a multilayer network where each node is a neuron and the edges in each layer would represent (a) structural connectivity, (b) functional connectivity, or (c) synapse types and numbers. These multilayer networks offer an attractive option to integrate multiple types of data, as reviewed in Boccaletti et al. (2014) and discussed later. Additionally, sheaves are a mathematical framework that allows one to represent not only the connections between nodes (as in the networks above), but also the specific relationship, such as linear maps, between node activities or properties that these connections represent. Thus far sheaves have been used for network coding problems (Ghrist & Hiraoka, 2011), signal processing (Robinson, 2013), and other applications (Phan-Luong, 2008), but they have great potential for use in biological research (Blevins & Bassett, 2020; Hansen & Ghrist, 2020). Though certainly not complete, we offer the above examples to show the seemingly limitless possibilities of mathematical system representations (Battiston et al., 2020; Torres et al., 2020).

By encoding data as a set of pairwise interactions, a network, or a simplicial complex, investigators can use statistical and computational techniques to describe the patterns they see. However, description does not amount to explanation, and nor does a measurement automatically produce an understanding. To move beyond description and measurement, and toward explanation and understanding, it can often be useful to define generative models (Betzel & Bassett, 2017a). With what ingredients and assumptions could one generate the observed pairwise interactions? Or the network? Or the simplicial complex? If one were able to devise a set of rules that produced data similar to those observed, then one would have identified a candidate mechanism for the observations (Bertolero & Bassett, 2020). Generative models do just that. A generative model is a set of interaction rules that produces an outcome; and hence, generative modeling as an approach can be used to determine what candidate mechanisms may have produced the observed data. Although generative models exist at all scales of analysis, here we will focus on generative models for networks. For those readers interested in other scales, see Kanashiro et al. (2017) for a cellular-scale generative model of interactions and Sizemore Blevins and Bassett (2020) for generative models of simplicial complexes.

Generative network models are defined by the specification of a wiring rule. Typically, wiring rules are defined probabilistically. For example, a node has a probability of connecting to other nodes with certain features. In brain networks, two features are commonly evaluated: (a) a geometric feature such as the physical distance between two nodes, and (b) a topological feature such as the degree or clustering coefficient of a node. For example, one might generate a network in which the probability of connecting node i to node j is negatively associated with the distance between them and positively associated with the product of the degree of node i and the degree of node j. The geometric and topological factors formally encode the two competing drivers of connectivity: wiring minimization and topological efficiency (Bullmore & Sporns, 2012). A minimally wired network, in which only the shortest possible links are present, would contain only edges between nearby nodes, requiring inefficient long relays to get information from one part of the network to another. Conversely, a maximally efficient topology, one in which topological path lengths are short, would contain edges between randomly chosen pairs of nodes, requiring costly long-distance tracts. By balancing these two competing pressures, the two-feature generative models are able to produce networks with topologies that are notably similar to those observed in real neural systems.

Generative network models are fairly young, and therefore a relatively small literature exists applying them to open questions in systems neuroscience. Yet despite their youth, generative network models have already been considered across species from C. elegans (Nicosia et al., 2013) to humans (Betzel et al., 2016), from cellular connectomes (Vértes et al., 2014) to macroscale connectomes (Betzel et al., 2016), and from healthy states to diseases states (Vértes et al., 2012; X. Zhang et al., 2021). Efforts have begun extending generative network models from the probabilistic so-called single-shot models that we have just described to growth models; these extensions pose the question of what growth rule (rather than what static wiring rule) might explain the observed trajectories of development, aging, or disease progression. A particularly exciting endeavor along these lines is the examination of a network morphospace constructed by the differential or combined expression of N network features (Avena-Koenigsberger et al., 2014; Avena-Koenigsberger et al., 2015). The approach allows the investigator to define a given network starting point (e.g., an infant connectome, or a healthy connectome) and rewire it to maximize or minimize a predefined set of network features, using a multi-objective evolutionary algorithm. The goal is to ask whether that rewiring rule can take the network to a final target point (e.g., a young adult connectome, or a connectome characteristic of a decade or more of neurodegeneration), precisely along the trajectory shown by development or by disease progression (Tang et al., 2017). Such efforts offer the exciting potential to test network hypotheses in a much more formal way than was hitherto possible, and to either validate or disprove posited mechanisms.

The methods described above focus on the functional (in the sense of the inferred function of the neurons) and temporal data from calcium imaging, but an additional challenge will be to integrate those analyses with the multimodal information that can be collected in parallel with calcium imaging. The additional modalities can be the physical position of the neurons (Betzel, 2020; X. Chen et al., 2018; Chiang et al., 2011), their ultrastructure (Hildebrand et al., 2017; Vishwanathan et al., 2022; Zheng et al., 2018), molecular subtype (Lovett-Barron et al., 2017; Shainer et al., 2022), morphology (Kunst et al., 2019), or even other cell-signaling reporters (Linghu et al., 2020). The data from these various modalities can be categorical (subtypes of neurons, morphology), time series (other cell signaling reporters), or numerical values (position and projection trajectory), each potentially coming with their own metrics and specific questions. The Euclidean distance between neurons’ spatial locations is easy to quantify, for example, but it is less easy to know how to integrate these types of data, alongside functional data, in the networks, or hypergraphs, and how to weight each modality to construct the model.

Despite the difficulty of the question of how to integrate multimodal datasets, researchers from many disciplines have pushed forward in this direction. First, multilayer frameworks offer one solution to incorporating multimodal data into one quantitative object. For example, multilayer networks can combine multiple types of neuron connectivity (Schroeder et al., 2022), activity from multiple frequency bands (Yu et al., 2017), or even data across individuals (Vaiana & Muldoon, 2020) within one object. Notably, the multilayer approach is not restricted to networks. Indeed, one could encode data as a multilayer simplicial complex or hypergraph following a similar procedure. Alternatively, we can incorporate data about each node using a node-annotated version of the above frameworks. In an annotated or attributed network, each node is also associated with a category (Bothorel et al., 2015; Yang et al., 2013) or a continuous value. These extra data are then incorporated into analytical methods, for example in performing community detection using task-based fMRI data (Murphy et al., 2016). Extending these ideas naturally leads one to matrix-valued graphs (Trinh et al., 2018) and sheaves (Robinson, 2017), which incorporate additional quantitative information on the relations between nodes and node activity. Still, as we previously discussed, no one framework fits every biological system and experiment, so we leave the finding of the perfect framework as an open question, as there is still incredible space for creative solutions.

Though much of this work has focused on systems composed of individual neurons, an additional future direction might consider how complex systems of neurons connect to that of the entire brain (Betzel & Bassett, 2017b). For example, networks could be created from data collected at the neuron level, population level, and the whole-brain level. In order to handle these multiple scales, one approach would create one network including all data, and use hierarchical methods or community detection to find subnetwork structures and determine how they fit together (Fortunato, 2010; Meunier et al., 2009). Alternatively, one could create a multilayer network with layers corresponding to the different neural scales (Betzel & Bassett, 2017b). Such a multilayer approach could even incorporate multiscale temporal data as well (Betzel & Bassett, 2017b). Networks of networks are even possible and often studied under the name of interdependent networks using a multilayer network framework (Gao et al., 2011). In TDA, the mathematics behind the methods allow for simple deconstruction of topological motifs of small complexes that are sub-complexes of a larger simplicial complex, thus allowing these tools to naturally incorporate multiscale data (Yoon & Ghrist, 2020). Finally, at a more fundamental level one can control the resolution using the definition of nodes. For example, one might define a node based on one neuron, a group of neurons, or even similarly clustered data (Betzel, 2020; van Veen et al., 2019).

As described above, graphs and network science may provide a common framework in which to integrate multiple information modalities (function, position, and neuronal identity, among others), and multiple scales of information. Notably, graphs can incorporate changes across time and represent the changing dynamics of the brain in different circumstances. The visualization and analytical tools offered by network science can look beyond pairwise interactions and reveal larger patterns of interactions or dynamics. As such, graphs provide a powerful basis on which to generate models of information processing and flow. Based on the observed network statistics, one can test a set of wiring rules and evaluate whether it generates a model similar to the observed network. Growth rules can be added to modify the wiring of the model across various trajectories such as development or disease. These rules can provide testable hypotheses on the mechanism underlying the observed dynamics of the network. Overall, the best test of any model is in its prediction accuracy, combined with advances in causal inferences, and circuit disruption technologies. One can imagine a scenario where you can predict how your model would react to a given perturbation of its wiring, and test it in vivo.

Finally, while this review focuses on systems neuroscience, these methods can be used in any quantitative biological context. In ecology, one can use causal inference to identify how species affect each other (Sugihara et al., 2012; D. D. Zhang et al., 2011) or network theory approaches to model food webs or spatial interactions (Casselberry et al., 2020; Krause et al., 2003). The same applies to cell biology, which benefited immensely from network approaches for characterizing regulatory gene networks and protein-protein interactions, among other interactions (Aittokallio & Schwikowski, 2006; Horvath & Dong, 2008; Martin et al., 2010). With the expansion of calcium imaging into nonneural tissues, the tools discussed here will also provide additional opportunities to better quantify, model, and ultimately understand the role calcium and intercell communication may play in diverse developmental and physiological processes (Balaji et al., 2017; A. J. Stevenson et al., 2020).

The availability of these types of data and their compatibility and searchability, as well as innovative visualization techniques, will all influence how easily the integration can occur (Akil et al., 2011). Efforts such as those led by the International Neuroinformatics Coordinating Facility (BRAINS Expert Working Group et al., 2017; Eglen et al., 2017) exist to standardize and unify neuroscience datasets, but they have not yet been embraced by the community at large, and we still need to ensure the code can be reusable (Riquelme & Gjorgjieva, 2021). Such commitments and infrastructural resources could serve to increase the power of the analysis of such data, and help to ensure the reproducibility of the research.

We have compiled and explained several of the most common data analysis approaches available to systems neuroscientists. As we show, no single method is superior to all the others and several are complimentary; researchers need to understand each method’s limitations to choose the most appropriate one for their specific question. Knowing which analysis will be used, and its limitations, can also guide experimental design to ensure that the chosen method can be used to its full potential on the dataset. For example, tools that can uncover high-dimensional links, such as hypergraphs, would allow for more complex stimuli or behaviors to be tested. On the contrary, they may require more parameter tuning, or have more built-in assumptions, which may generate spurious results with low-dimensional experiments. We also suggest that the complex systems field may provide robust frameworks with which to integrate multiple streams of information. For example, multilayer networks, or networks of networks, could represent different information modalities (functional imaging, molecular identity, physical location, or others) or different scales (subcellular, single cell, populations, or whole brain). These frameworks also benefit from the numerous tools that the wider mathematics community has developed. In conclusion, we propose that multilayer networks are our best tool to integrate multimodal information into a single framework.

Recent work in several fields of science has identified a bias in citation practices such that papers from women and other minority scholars are under-cited relative to the number of such papers in the field (Caplar et al., 2017; Dion et al., 2018; Dworkin et al., 2020; Maliniak et al., 2013; Mitchell et al., 2013). Here we sought to proactively consider choosing references that reflect the diversity of the field in thought, form of contribution, gender, race, ethnicity, and other factors. First, we obtained the predicted gender of the first and last author of each reference by using databases that store the probability of a first name being carried by a woman (Dworkin et al., 2020; Zhou et al., 2020). By this measure (and excluding self-citations to the first and last authors of our current paper), our references contain 7.05% woman (first) / woman (last), 7.68% man/woman, 15.18% woman/man, and 70.1% man/man. This method is limited in that (a) names, pronouns, and social media profiles used to construct the databases may not, in every case, be indicative of gender identity and (b) it cannot account for intersex, nonbinary, or transgender people. Second, we obtained the predicted racial/ethnic category of the first and last author of each reference by databases that store the probability of a first and last name being carried by an author of color (Ambekar et al., 2009; Sood & Laohaprapanon, 2018). By this measure (and excluding self-citations), our references contain 12.61% author of color (first) / author of color (last), 16.13% white author / author of color, 22.15% author of color / white author, and 49.11% white author / white author. This method is limited in that (a) names and Florida Voter Data to make the predictions may not be indicative of racial/ethnic identity, and (b) it cannot account for Indigenous and mixed-race authors, or those who may face differential biases due to the ambiguous racialization or ethnicization of their names. We look forward to future work that could help us to better understand how to support equitable practices in science.

Ann S. Blevins: Conceptualization; Visualization; Writing – original draft; Writing – review & editing. Dani S. Bassett: Conceptualization; Supervision; Writing – review & editing. Ethan K. Scott: Supervision; Writing – review & editing. Gilles Claude Vanwalleghem: Conceptualization; Visualization; Writing – original draft; Writing – review & editing.

DANDRITE, Lundbeckfonden (https://dx.doi.org/10.13039/501100003554), Award ID: DANDRITE-R248-2016-2518. Gilles Claude Vanwalleghem, Novo Nordisk Fonden (https://dx.doi.org/10.13039/501100009708). Gilles Claude Vanwalleghem, Aarhus Universitets Forskningsfond (https://dx.doi.org/10.13039/501100002739). Ethan K. Scott, National Health and Medical Research Council (https://dx.doi.org/10.13039/501100000925), Award ID: APP1066887. Ethan K. Scott, Australian Research Council (https://dx.doi.org/10.13039/501100000923), Award ID: DP140102036 - DP110103612. Ethan K. Scott, Australian Research Council (https://dx.doi.org/10.13039/501100000923), Award ID: FT110100887. Dani S. Bassett, Army Research Office (https://dx.doi.org/10.13039/100000183), Award ID: Bassett-W911NF-14-1-0679 and Grafton-W911NF-16-1-0474.

Calcium indicators:

Fluorescent proteins or dyes whose fluorescence changes with the calcium concentration, typically used as a proxy for neuronal activity.

Lateral line:

Sensory organ found in fish and amphibians; it is composed of neuromasts, specialized bundles of hair cells that can detect waterflow.

Optogenetics:

Genetic tools that allow the manipulation of cells by light, typically to activate or silence neurons.

Optic tectum:

Midbrain region that serves as the main visual processing center in several vertebrates, homologue to the superior colliculus of mammals.

Edge:

Connection between two nodes; it indicates they are related, and it can be weighted (measuring the strength of that relation) or unweighted (no information on the strength). It can also have direction or not, depending on the relation metric used.

Degree:

Number of connections (edges) from a node to other nodes (and itself).

Fragile X syndrome:

Genetic condition caused by the silencing of the FMR1 gene on the X chromosome (hence the name). The syndrome is linked to intellectual disability in humans.

Traumatic brain injury:

Damage to the brain from physical trauma, typically from a blow or impact to the head.

Multilayer networks:

These networks allow the interactions of nodes between different layers, each representing different node types, potentially representing different data types.

Ahissar
,
E.
, &
Kleinfeld
,
D.
(
2003
).
Closed-loop neuronal computations: Focus on vibrissa somatosensation in rat
.
Cerebral Cortex
,
13
(
1
),
53
62
. ,
[PubMed]
Aimon
,
S.
,
Katsuki
,
T.
,
Jia
,
T.
,
Grosenick
,
L.
,
Broxton
,
M.
,
Deisseroth
,
K.
,
Sejnowski
,
T. J.
, &
Greenspan
,
R. J.
(
2019
).
Fast near-whole-brain imaging in adult Drosophila during responses to stimuli and behavior
.
PLoS Biology
,
17
(
2
),
e2006732
. ,
[PubMed]
Aittokallio
,
T.
, &
Schwikowski
,
B.
(
2006
).
Graph-based methods for analysing networks in cell biology
.
Briefings in Bioinformatics
,
7
(
3
),
243
255
. ,
[PubMed]
Akil
,
H.
,
Martone
,
M. E.
, &
Van Essen
,
D. C.
(
2011
).
Challenges and opportunities in mining neuroscience data
.
Science
,
331
(
6018
),
708
712
. ,
[PubMed]
Ambekar
,
A.
,
Ward
,
C.
,
Mohammed
,
J.
,
Male
,
S.
, &
Skiena
,
S.
(
2009
).
Name-ethnicity classification from open sources
. In
Proceedings of the 15th ACM SIGKDD international conference on knowledge discovery and data mining - KDD ’09
.
Avena-Koenigsberger
,
A.
,
Goñi
,
J.
,
Betzel
,
R. F.
,
van den Heuvel
,
M. P.
,
Griffa
,
A.
,
Hagmann
,
P.
,
Thiran
,
J.-P.
, &
Sporns
,
O.
(
2014
).
Using Pareto optimality to explore the topology and dynamics of the human connectome
.
Philosophical Transactions of the Royal Society of London B: Biological Sciences
,
369
(
1653
),
20130530
. ,
[PubMed]
Avena-Koenigsberger
,
A.
,
Goñi
,
J.
,
Solé
,
R.
, &
Sporns
,
O.
(
2015
).
Network morphospace
.
Journal of the Royal Society Interface
,
12
(
103
),
20140881
. ,
[PubMed]
Avitan
,
L.
,
Pujic
,
Z.
,
Mölter
,
J.
,
McCullough
,
M.
,
Zhu
,
S.
,
Sun
,
B.
,
Myhre
,
A.-E.
, &
Goodhill
,
G. J.
(
2020
).
Behavioral signatures of a developing neural code
.
Current Biology
,
30
(
17
),
3352
3363
. ,
[PubMed]
Avitan
,
L.
,
Pujic
,
Z.
,
Mölter
,
J.
,
van de Poll
,
M.
,
Sun
,
B.
,
Teng
,
H.
,
Amor
,
R.
,
Scott
,
E. K.
, &
Goodhill
,
G. J.
(
2017
).
Spontaneous activity in the zebrafish tectum reorganizes over development and is influenced by visual experience
.
Current Biology
,
27
(
16
),
2407
2419
. ,
[PubMed]
Balaji
,
R.
,
Bielmeier
,
C.
,
Harz
,
H.
,
Bates
,
J.
,
Stadler
,
C.
,
Hildebrand
,
A.
, &
Classen
,
A.-K.
(
2017
).
Calcium spikes, waves and oscillations in a large, patterned epithelial tissue
.
Scientific Reports
,
7
,
42786
. ,
[PubMed]
Bardin
,
J.-B.
,
Spreemann
,
G.
, &
Hess
,
K.
(
2019
).
Topological exploration of artificial neuronal network dynamics
.
Network Neuroscience
,
3
(
3
),
725
743
. ,
[PubMed]
Barlow
,
H. B.
(
1972
).
Single units and sensation: A neuron doctrine for perceptual psychology? [Review of Single units and sensation: A neuron doctrine for perceptual psychology?]
.
Perception
,
1
(
4
),
371
394
. ,
[PubMed]
Bassett
,
D. S.
, &
Bullmore
,
E.
(
2006
).
Small-world brain networks
.
The Neuroscientist
,
12
(
6
),
512
523
. ,
[PubMed]
Bassett
,
D. S.
, &
Bullmore
,
E. T.
(
2017
).
Small-world brain networks revisited
.
The Neuroscientist
,
23
(
5
),
499
516
. ,
[PubMed]
Bassett
,
D. S.
, &
Mattar
,
M. G.
(
2017
).
A network neuroscience of human learning: Potential to inform quantitative theories of brain and behavior
.
Trends in Cognitive Sciences
,
21
(
4
),
250
264
. ,
[PubMed]
Bassett
,
D. S.
, &
Sporns
,
O.
(
2017
).
Network neuroscience
.
Nature Neuroscience
,
20
(
3
),
353
364
. ,
[PubMed]
Bassett
,
D. S.
,
Zurn
,
P.
, &
Gold
,
J. I.
(
2018
).
On the nature and use of models in network neuroscience
.
Nature Reviews Neuroscience
,
19
(
9
),
566
578
. ,
[PubMed]
Basso
,
E.
,
Arai
,
M.
, &
Dabaghian
,
Y.
(
2016
).
Gamma synchronization influences map formation time in a topological model of spatial learning
.
PLoS Computational Biology
,
12
(
9
),
e1005114
. ,
[PubMed]
Battiston
,
F.
,
Cencetti
,
G.
,
Iacopini
,
I.
,
Latora
,
V.
,
Lucas
,
M.
,
Patania
,
A.
,
Young
,
J.-G.
, &
Petri
,
G.
(
2020
).
Networks beyond pairwise interactions: Structure and dynamics
.
Physics Reports
,
874
,
1
92
.
Belilovsky
,
E.
,
Varoquaux
,
G.
, &
Blaschko
,
M. B.
(
2015
).
Testing for differences in Gaussian graphical models: Applications to brain connectivity
.
arXiv:1512.08643
.
Bertolero
,
M. A.
, &
Bassett
,
D. S.
(
2020
).
On the nature of explanations offered by network science: A perspective from and for practicing neuroscientists
.
Topics in Cognitive Science
,
12
(
4
),
1272
1293
. ,
[PubMed]
Betzel
,
R. F.
(
2020
).
Organizing principles of whole-brain functional connectivity in zebrafish larvae
.
Network Neuroscience
,
4
(
1
),
234
256
. ,
[PubMed]
Betzel
,
R. F.
,
Avena-Koenigsberger
,
A.
,
Goñi
,
J.
,
He
,
Y.
,
de Reus
,
M. A.
,
Griffa
,
A.
,
Vértes
,
P. E.
,
Mišic
,
B.
,
Thiran
,
J.-P.
,
Hagmann
,
P.
,
van den Heuvel
,
M.
,
Zuo
,
X.-N.
,
Bullmore
,
E. T.
, &
Sporns
,
O.
(
2016
).
Generative models of the human connectome
.
NeuroImage
,
124
(
Pt. A
),
1054
1064
. ,
[PubMed]
Betzel
,
R. F.
, &
Bassett
,
D. S.
(
2017a
).
Generative models for network neuroscience: Prospects and promise
.
Journal of the Royal Society Interface
,
14
(
136
),
20170623
. ,
[PubMed]
Betzel
,
R. F.
, &
Bassett
,
D. S.
(
2017b
).
Multi-scale brain networks
.
NeuroImage
,
160
,
73
83
. ,
[PubMed]
Betzel
,
R. F.
, &
Bassett
,
D. S.
(
2018
).
Specificity and robustness of long-distance connections in weighted, interareal connectomes
.
Proceedings of the National Academy of Sciences
,
115
(
21
),
E4880
E4889
. ,
[PubMed]
Bielza
,
C.
, &
Larrañaga
,
P.
(
2014
).
Bayesian networks in neuroscience: A survey
.
Frontiers in Computational Neuroscience
,
8
,
131
. ,
[PubMed]
Blevins
,
A. S.
, &
Bassett
,
D. S.
(
2020
).
Topology in biology
. In
Handbook of the Mathematics of the Arts and Sciences
(pp.
1
23
).
Bobadilla-Suarez
,
S.
,
Ahlheim
,
C.
,
Mehrotra
,
A.
,
Panos
,
A.
, &
Love
,
B. C.
(
2020
).
Measures of neural similarity
.
Computational Brain and Behavior
,
3
(
4
),
369
383
. ,
[PubMed]
Boccaletti
,
S.
,
Bianconi
,
G.
,
Criado
,
R.
,
Del Genio
,
C. I.
,
Gómez-Gardeñes
,
J.
,
Romance
,
M.
,
Sendiña-Nadal
,
I.
,
Wang
,
Z.
, &
Zanin
,
M.
(
2014
).
The structure and dynamics of multilayer networks
.
Physics Reports
,
544
(
1
),
1
122
. ,
[PubMed]
Bothorel
,
C.
,
Cruz
,
J. D.
,
Magnani
,
M.
, &
Micenková
,
B.
(
2015
).
Clustering attributed graphs: Models, measures and methods
.
Network Science
,
3
(
3
),
408
444
.
BRAINS (Brain Imaging in Normal Subjects) Expert Working Group
,
Shenkin
,
S. D.
,
Pernet
,
C.
,
Nichols
,
T. E.
,
Poline
,
J.-B.
,
Matthews
,
P. M.
,
van der Lugt
,
A.
,
Mackay
,
C.
,
Lanyon
,
L.
,
Mazoyer
,
B.
,
Boardman
,
J. P.
,
Thompson
,
P. M.
,
Fox
,
N.
,
Marcus
,
D. S.
,
Sheikh
,
A.
,
Cox
,
S. R.
,
Anblagan
,
D.
,
Job
,
D. E.
,
Dickie
,
D. A.
, …
Wardlaw
,
J. M.
(
2017
).
Improving data availability for brain image biobanking in healthy subjects: Practice-based suggestions from an international multidisciplinary working group
.
NeuroImage
,
153
,
399
409
. ,
[PubMed]
Branner
,
A.
, &
Normann
,
R. A.
(
2000
).
A multielectrode array for intrafascicular recording and stimulation in sciatic nerve of cats
.
Brain Research Bulletin
,
51
(
4
),
293
306
. ,
[PubMed]
Broido
,
A. D.
, &
Clauset
,
A.
(
2019
).
Scale-free networks are rare
.
Nature Communications
,
10
(
1
),
1017
. ,
[PubMed]
Buldú
,
J. M.
, &
Porter
,
M. A.
(
2018
).
Frequency-based brain networks: From a multiplex framework to a full multilayer description
.
Network Neuroscience
,
2
(
4
),
418
441
. ,
[PubMed]
Bullmore
,
E.
, &
Sporns
,
O.
(
2009
).
Complex brain networks: Graph theoretical analysis of structural and functional systems
.
Nature Reviews Neuroscience
,
10
(
3
),
186
198
. ,
[PubMed]
Bullmore
,
E.
, &
Sporns
,
O.
(
2012
).
The economy of brain network organization
.
Nature Reviews Neuroscience
,
13
(
5
),
336
349
. ,
[PubMed]
Cámara
,
P. G.
(
2017
).
Topological methods for genomics: Present and future directions
.
Current Opinion in Systems Biology
,
1
,
95
101
. ,
[PubMed]
Caplar
,
N.
,
Tacchella
,
S.
, &
Birrer
,
S.
(
2017
).
Quantitative evaluation of gender bias in astronomical publications from citation counts
.
Nature Astronomy
,
1
(
6
),
0141
.
Carrillo-Reid
,
L.
,
Han
,
S.
,
O’Neil
,
D.
,
Taralova
,
E.
,
Jebara
,
T.
, &
Yuste
,
R.
(
2021
).
Identification of pattern completion neurons in neuronal ensembles using probabilistic graphical models
.
Journal of Neuroscience
,
41
(
41
),
8577
8588
. ,
[PubMed]
Casselberry
,
G. A.
,
Danylchuk
,
A. J.
,
Finn
,
J. T.
,
DeAngelis
,
B. M.
,
Jordaan
,
A.
,
Pollock
,
C. G.
,
Lundgren
,
I.
,
Hillis-Starr
,
Z.
, &
Skomal
,
G. B.
(
2020
).
Network analysis reveals multispecies spatial associations in the shark community of a Caribbean marine protected area
.
Marine Ecology Progress Series
,
633
,
105
126
.
Cekic
,
S.
,
Grandjean
,
D.
, &
Renaud
,
O.
(
2018
).
Time, frequency, and time-varying Granger-causality measures in neuroscience
.
Statistics in Medicine
,
37
(
11
),
1910
1931
. ,
[PubMed]
Chang
,
A.
,
Yao
,
T.
, &
Allen
,
G. I.
(
2019
).
Graphical models and dynamic latent factors for modeling functional brain connectivity
.
2019 IEEE Data Science Workshop (DSW)
,
57
63
.
Chen
,
T.-W.
,
Wardill
,
T. J.
,
Sun
,
Y.
,
Pulver
,
S. R.
,
Renninger
,
S. L.
,
Baohan
,
A.
,
Schreiter
,
E. R.
,
Kerr
,
R. A.
,
Orger
,
M. B.
,
Jayaraman
,
V.
,
Looger
,
L. L.
,
Svoboda
,
K.
, &
Kim
,
D. S.
(
2013
).
Ultrasensitive fluorescent proteins for imaging neuronal activity
.
Nature
,
499
(
7458
),
295
300
. ,
[PubMed]
Chen
,
X.
,
Mu
,
Y.
,
Hu
,
Y.
,
Kuan
,
A. T.
,
Nikitchenko
,
M.
,
Randlett
,
O.
,
Chen
,
A. B.
,
Gavornik
,
J. P.
,
Sompolinsky
,
H.
,
Engert
,
F.
, &
Ahrens
,
M. B.
(
2018
).
Brain-wide organization of neuronal activity and convergent sensorimotor transformations in larval zebrafish
.
Neuron
,
100
(
4
),
876
890
. ,
[PubMed]
Chiang
,
A.-S.
,
Lin
,
C.-Y.
,
Chuang
,
C.-C.
,
Chang
,
H.-M.
,
Hsieh
,
C.-H.
,
Yeh
,
C.-W.
,
Shih
,
C.-T.
,
Wu
,
J.-J.
,
Wang
,
G.-T.
,
Chen
,
Y.-C.
,
Wu
,
C.-C.
,
Chen
,
G.-Y.
,
Ching
,
Y.-T.
,
Lee
,
P.-C.
,
Lin
,
C.-Y.
,
Lin
,
H.-H.
,
Wu
,
C.-C.
,
Hsu
,
H.-W.
,
Huang
,
Y.-A.
, …
Hwang
,
J.-K.
(
2011
).
Three-dimensional reconstruction of brain-wide wiring networks in Drosophila at single-cell resolution
.
Current Biology
,
21
(
1
),
1
11
. ,
[PubMed]
Cliff
,
O. M.
,
Lizier
,
J. T.
,
Tsuchiya
,
N.
, &
Fulcher
,
B. D.
(
2022
).
Unifying pairwise interactions in complex dynamics
.
arXiv:2201.11941
.
Constantin
,
L.
,
Poulsen
,
R. E.
,
Scholz
,
L. A.
,
Favre-Bulle
,
I. A.
,
Taylor
,
M. A.
,
Sun
,
B.
,
Goodhill
,
G. J.
,
Vanwalleghem
,
G. C.
, &
Scott
,
E. K.
(
2020
).
Altered brain-wide auditory networks in a zebrafish model of fragile X syndrome
.
BMC Biology
,
18
(
1
),
125
. ,
[PubMed]
Crossley
,
N. A.
,
Mechelli
,
A.
,
Vértes
,
P. E.
,
Winton-Brown
,
T. T.
,
Patel
,
A. X.
,
Ginestet
,
C. E.
,
McGuire
,
P.
, &
Bullmore
,
E. T.
(
2013
).
Cognitive relevance of the community structure of the human brain functional coactivation network
.
Proceedings of the National Academy of Sciences
,
110
(
28
),
11583
11588
. ,
[PubMed]
Curto
,
C.
(
2016
).
What can topology tell us about the neural code?
Bulletin of the American Mathematical Society
,
54
(
1
),
63
78
.
Curto
,
C.
,
Degeratu
,
A.
, &
Itskov
,
V.
(
2012
).
Flexible memory networks
.
Bulletin of Mathematical Biology
,
74
(
3
),
590
614
. ,
[PubMed]
Cutts
,
C. S.
, &
Eglen
,
S. J.
(
2014
).
Detecting pairwise correlations in spike trains: An objective comparison of methods and application to the study of retinal waves
.
Journal of Neuroscience
,
34
(
43
),
14288
14303
. ,
[PubMed]
da Fontoura Costa
,
L.
,
Rodrigues
,
F. A.
,
Travieso
,
G.
, &
Villas Boas
,
P. R.
(
2007
).
Characterization of complex networks: A survey of measurements
.
Advances in Physics
,
56
(
1
),
167
242
.
Dal Maschio
,
M.
,
Donovan
,
J. C.
,
Helmbrecht
,
T. O.
, &
Baier
,
H.
(
2017
).
Linking neurons to network function and behavior by two-photon holographic optogenetics and volumetric imaging
.
Neuron
,
94
(
4
),
774
789
. ,
[PubMed]
Davison
,
E. N.
,
Schlesinger
,
K. J.
,
Bassett
,
D. S.
,
Lynall
,
M.-E.
,
Miller
,
M. B.
,
Grafton
,
S. T.
, &
Carlson
,
J. M.
(
2015
).
Brain network adaptability across task states
.
PLoS Computational Biology
,
11
(
1
),
e1004029
. ,
[PubMed]
Davison
,
E. N.
,
Turner
,
B. O.
,
Schlesinger
,
K. J.
,
Miller
,
M. B.
,
Grafton
,
S. T.
,
Bassett
,
D. S.
, &
Carlson
,
J. M.
(
2016
).
Individual differences in dynamic functional brain connectivity across the human lifespan
.
PLoS Computational Biology
,
12
(
11
),
e1005178
. ,
[PubMed]
Dimitriadis
,
S. I.
,
Lancaster
,
T. M.
,
Perry
,
G.
,
Tansey
,
K. E.
,
Jones
,
D. K.
,
Singh
,
K. D.
,
Zammit
,
S.
,
Davey Smith
,
G.
,
Hall
,
J.
,
O’Donovan
,
M. C.
,
Owen
,
M. J.
, &
Linden
,
D. E.
(
2021
).
Global brain flexibility during working memory is reduced in a high genetic risk group for schizophrenia
.
Biological Psychiatry: Cognitive Neuroscience and Neuroimaging
,
6
(
12
),
1176
1184
. ,
[PubMed]
Dion
,
M. L.
,
Sumner
,
J. L.
, &
Mitchell
,
S. M.
(
2018
).
Gendered citation patterns across political science and social science methodology fields
.
Political Analysis
,
26
(
3
),
312
327
.
Dunn
,
T. W.
,
Mu
,
Y.
,
Narayan
,
S.
,
Randlett
,
O.
,
Naumann
,
E. A.
,
Yang
,
C.-T.
,
Schier
,
A. F.
,
Freeman
,
J.
,
Engert
,
F.
, &
Ahrens
,
M. B.
(
2016
).
Brain-wide mapping of neural activity controlling zebrafish exploratory locomotion
.
eLife
,
5
,
e12741
. ,
[PubMed]
Dworkin
,
J. D.
,
Linn
,
K. A.
,
Teich
,
E. G.
,
Zurn
,
P.
,
Shinohara
,
R. T.
, &
Bassett
,
D. S.
(
2020
).
The extent and drivers of gender imbalance in neuroscience reference lists
.
Nature Neuroscience
,
23
(
8
),
918
926
. ,
[PubMed]
Eglen
,
S. J.
,
Marwick
,
B.
,
Halchenko
,
Y. O.
,
Hanke
,
M.
,
Sufi
,
S.
,
Gleeson
,
P.
,
Silver
,
R. A.
,
Davison
,
A. P.
,
Lanyon
,
L.
,
Abrams
,
M.
,
Wachtler
,
T.
,
Willshaw
,
D. J.
,
Pouzat
,
C.
, &
Poline
,
J.-B.
(
2017
).
Toward standard practices for sharing computer code and programs in neuroscience
.
Nature Neuroscience
,
20
(
6
),
770
773
. ,
[PubMed]
Fallani
,
F. D. V.
,
Corazzol
,
M.
,
Sternberg
,
J. R.
,
Wyart
,
C.
, &
Chavez
,
M.
(
2015
).
Hierarchy of neural organization in the embryonic spinal cord: Granger-causality graph analysis of in vivo calcium imaging data
.
IEEE Transactions on Neural Systems and Rehabilitation Engineering
,
23
(
3
),
333
341
. ,
[PubMed]
Favre-Bulle
,
I. A.
,
Taylor
,
M. A.
,
Marquez-Legorreta
,
E.
,
Vanwalleghem
,
G.
,
Poulsen
,
R. E.
,
Rubinsztein-Dunlop
,
H.
, &
Scott
,
E. K.
(
2020
).
Sound generation in zebrafish with Bio-Opto-Acoustics
.
Nature Communications
,
11
(
1
),
6120
. ,
[PubMed]
Favre-Bulle
,
I. A.
,
Vanwalleghem
,
G.
,
Taylor
,
M. A.
,
Rubinsztein-Dunlop
,
H.
, &
Scott
,
E. K.
(
2018
).
Cellular-resolution imaging of vestibular processing across the larval zebrafish brain
.
Current Biology
,
28
(
23
),
3711
3722
. ,
[PubMed]
Fernandes
,
A. M.
,
Mearns
,
D. S.
,
Donovan
,
J. C.
,
Larsch
,
J.
,
Helmbrecht
,
T. O.
,
Kölsch
,
Y.
,
Laurell
,
E.
,
Kawakami
,
K.
,
Dal Maschio
,
M.
, &
Baier
,
H.
(
2021
).
Neural circuitry for stimulus selection in the zebrafish visual system
.
Neuron
,
109
(
5
),
805
822
. ,
[PubMed]
Fornito
,
A.
,
Zalesky
,
A.
, &
Bullmore
,
E.
(
2016
).
Fundamentals of brain network analysis
.
Academic Press
.
Fortunato
,
S.
(
2010
).
Community detection in graphs
.
Physics Reports
,
486
(
3
),
75
174
.
Freeman
,
J.
,
Vladimirov
,
N.
,
Kawashima
,
T.
,
Mu
,
Y.
,
Sofroniew
,
N. J.
,
Bennett
,
D. V.
,
Rosen
,
J.
,
Yang
,
C.-T.
,
Looger
,
L. L.
, &
Ahrens
,
M. B.
(
2014
).
Mapping brain activity at scale with cluster computing
.
Nature Methods
,
11
(
9
),
941
950
. ,
[PubMed]
Friedrich
,
R. W.
,
Habermann
,
C. J.
, &
Laurent
,
G.
(
2004
).
Multiplexing using synchrony in the zebrafish olfactory bulb
.
Nature Neuroscience
,
7
(
8
),
862
871
. ,
[PubMed]
Gao
,
J.
,
Buldyrev
,
S. V.
,
Eugene Stanley
,
H.
, &
Havlin
,
S.
(
2011
).
Networks formed from interdependent networks
.
Nature Physics
,
8
(
1
),
40
48
.
Geniesse
,
C.
,
Chowdhury
,
S.
, &
Saggar
,
M.
(
2022
).
NeuMapper: A scalable computational framework for multiscale exploration of the brain’s dynamical organization
.
Network Neuroscience
,
6
(
2
),
467
498
. ,
[PubMed]
Geniesse
,
C.
,
Sporns
,
O.
,
Petri
,
G.
, &
Saggar
,
M.
(
2019
).
Generating dynamical neuroimaging spatiotemporal representations (DyNeuSR) using topological data analysis
.
Network Neuroscience
,
3
(
3
),
763
778
. ,
[PubMed]
Ghrist
,
R.
(
2007
).
Barcodes: The persistent topology of data
.
Bulletin of the American Mathematical Society
,
45
(
1
),
61
76
.
Ghrist
,
R.
, &
Hiraoka
,
Y.
(
2011
).
Applications of sheaf cohomology and exact sequences on network codings
.
Preprint
. https://citeseerx.ist.psu.edu/viewdoc/download?doi=10.1.1.224.6433&rep=rep1&type=pdf
Giusti
,
C.
,
Pastalkova
,
E.
,
Curto
,
C.
, &
Itskov
,
V.
(
2015
).
Clique topology reveals intrinsic geometric structure in neural correlations
.
Proceedings of the National Academy of Sciences
,
112
(
44
),
13455
13460
. ,
[PubMed]
Gracia-Tabuenca
,
Z.
,
Díaz-Patiño
,
J. C.
,
Arelio
,
I.
, &
Alcauter
,
S.
(
2020
).
Topological data analysis reveals robust alterations in the whole-brain and frontal lobe functional connectomes in attention-deficit/hyperactivity disorder
.
bioRxiv
.
Granger
,
C. W. J.
(
1969
).
Investigating causal relations by econometric models and cross-spectral methods
.
Econometrica
,
37
(
3
),
424
438
.
Haesemeyer
,
M.
,
Schier
,
A. F.
, &
Engert
,
F.
(
2019
).
Convergent temperature representations in artificial and biological neural networks
.
Neuron
,
103
(
6
),
1123
1134
. ,
[PubMed]
Hansen
,
J.
, &
Ghrist
,
R.
(
2020
).
Opinion dynamics on discourse sheaves
.
arXiv:2005.12798
.
Hastings
,
J. W.
,
Mitchell
,
G.
,
Mattingly
,
P. H.
,
Blinks
,
J. R.
, &
Van Leeuwen
,
M.
(
1969
).
Response of aequorin bioluminescence to rapid changes in calcium concentration
.
Nature
,
222
(
5198
),
1047
1050
. ,
[PubMed]
Helmbrecht
,
T. O.
,
Dal Maschio
,
M.
,
Donovan
,
J. C.
,
Koutsouli
,
S.
, &
Baier
,
H.
(
2018
).
Topography of a visuomotor transformation
.
Neuron
,
100
(
6
),
1429
1445
. ,
[PubMed]
Hildebrand
,
D. G. C.
,
Cicconet
,
M.
,
Torres
,
R. M.
,
Choi
,
W.
,
Quan
,
T. M.
,
Moon
,
J.
,
Wetzel
,
A. W.
,
Scott Champion
,
A.
,
Graham
,
B. J.
,
Randlett
,
O.
,
Plummer
,
G. S.
,
Portugues
,
R.
,
Bianco
,
I. H.
,
Saalfeld
,
S.
,
Baden
,
A. D.
,
Lillaney
,
K.
,
Burns
,
R.
,
Vogelstein
,
J. T.
,
Schier
,
A. F.
, …
Engert
,
F.
(
2017
).
Whole-brain serial-section electron microscopy in larval zebrafish
.
Nature
,
545
(
7654
),
345
349
. ,
[PubMed]
Hilgetag
,
C. C.
, &
Goulas
,
A.
(
2016
).
Is the brain really a small-world network?
Brain Structure and Function
,
221
(
4
),
2361
2366
. ,
[PubMed]
Horvath
,
S.
, &
Dong
,
J.
(
2008
).
Geometric interpretation of gene coexpression network analysis
.
PLoS Computational Biology
,
4
(
8
),
e1000117
. ,
[PubMed]
Ince
,
R. A. A.
,
Panzeri
,
S.
, &
Kayser
,
C.
(
2013
).
Neural codes formed by small and temporally precise populations in auditory cortex
.
Journal of Neuroscience
,
33
(
46
),
18277
18287
. ,
[PubMed]
Javidian
,
M. A.
,
Wang
,
Z.
,
Lu
,
L.
, &
Valtorta
,
M.
(
2020
).
On a hypergraph probabilistic graphical model
.
Annals of Mathematics and Artificial Intelligence
,
88
(
9
),
1003
1033
.
Kaiser
,
M.
, &
Varier
,
S.
(
2011
).
Evolution and development of brain networks: From Caenorhabditis elegans to Homo sapiens
.
Network
,
22
(
1–4
),
143
147
. ,
[PubMed]
Kanashiro
,
T.
,
Ocker
,
G. K.
,
Cohen
,
M. R.
, &
Doiron
,
B.
(
2017
).
Attentional modulation of neuronal variability in circuit models of cortex
.
eLife
,
6
,
e23978
. ,
[PubMed]
Kim
,
D. H.
,
Kim
,
J.
,
Marques
,
J. C.
,
Grama
,
A.
,
Hildebrand
,
D. G. C.
,
Gu
,
W.
,
Li
,
J. M.
, &
Robson
,
D. N.
(
2017
).
Pan-neuronal calcium imaging with cellular resolution in freely swimming zebrafish
.
Nature Methods
,
14
(
11
),
1107
1114
. ,
[PubMed]
Knight
,
B. W.
(
1972
).
Dynamics of encoding in a population of neurons
.
Journal of General Physiology
,
59
(
6
),
734
766
. ,
[PubMed]
Krause
,
A. E.
,
Frank
,
K. A.
,
Mason
,
D. M.
,
Ulanowicz
,
R. E.
, &
Taylor
,
W. W.
(
2003
).
Compartments revealed in food-web structure
.
Nature
,
426
(
6964
),
282
285
. ,
[PubMed]
Kunst
,
M.
,
Laurell
,
E.
,
Mokayes
,
N.
,
Kramer
,
A.
,
Kubo
,
F.
,
Fernandes
,
A. M.
,
Förster
,
D.
,
Dal Maschio
,
M.
, &
Baier
,
H.
(
2019
).
A cellular-resolution atlas of the larval zebrafish brain
.
Neuron
,
103
(
1
),
21
38
. ,
[PubMed]
Lepperød
,
M. E.
,
Stöber
,
T.
,
Hafting
,
T.
,
Fyhn
,
M.
, &
Kording
,
K. P.
(
2018
).
Inferring causal connectivity from pairwise recordings and optogenetics
.
bioRxiv
.
Linghu
,
C.
,
Johnson
,
S. L.
,
Valdes
,
P. A.
,
Shemesh
,
O. A.
,
Park
,
W. M.
,
Park
,
D.
,
Piatkevich
,
K. D.
,
Wassie
,
A. T.
,
Liu
,
Y.
,
An
,
B.
,
Barnes
,
S. A.
,
Celiker
,
O. T.
,
Yao
,
C.-C.
,
Yu
,
C.-C. J.
,
Wang
,
R.
,
Adamala
,
K. P.
,
Bear
,
M. F.
,
Keating
,
A. E.
, &
Boyden
,
E. S.
(
2020
).
Spatial multiplexing of fluorescent reporters for imaging signaling network dynamics
.
Cell
,
183
(
6
),
1682
1698
. ,
[PubMed]
Lisman
,
J. E.
, &
Grace
,
A. A.
(
2005
).
The hippocampal-VTA loop: Controlling the entry of information into long-term memory
.
Neuron
,
46
(
5
),
703
713
. ,
[PubMed]
Lovett-Barron
,
M.
,
Andalman
,
A. S.
,
Allen
,
W. E.
,
Vesuna
,
S.
,
Kauvar
,
I.
,
Burns
,
V. M.
, &
Deisseroth
,
K.
(
2017
).
Ancestral circuits for the coordinated modulation of brain state
.
Cell
,
171
(
6
),
1411
1423
. ,
[PubMed]
Luchiari
,
A. C.
, &
Chacon
,
D. M. M.
(
2013
).
Physical exercise improves learning in zebrafish, Danio rerio
.
Behavioural Processes
,
100
,
44
47
. ,
[PubMed]
Ma’ayan
,
A.
,
Cecchi
,
G. A.
,
Wagner
,
J.
,
Rao
,
A. R.
,
Iyengar
,
R.
, &
Stolovitzky
,
G.
(
2008
).
Ordered cyclic motifs contribute to dynamic stability in biological and engineered networks
.
Proceedings of the National Academy of Sciences
,
105
(
49
),
19235
19240
. ,
[PubMed]
Maliniak
,
D.
,
Powers
,
R.
, &
Walter
,
B. F.
(
2013
).
The gender citation gap in international relations
.
International Organization
,
67
(
4
),
889
922
.
Marinescu
,
I. E.
,
Lawlor
,
P. N.
, &
Kording
,
K. P.
(
2018
).
Quasi-experimental causality in neuroscience and behavioural research
.
Nature Human Behaviour
,
2
(
12
),
891
898
. ,
[PubMed]
Marques
,
J. C.
,
Li
,
M.
,
Schaak
,
D.
,
Robson
,
D. N.
, &
Li
,
J. M.
(
2020
).
Internal state dynamics shape brainwide activity and foraging behaviour
.
Nature
,
577
(
7789
),
239
243
. ,
[PubMed]
Marshel
,
J. H.
,
Kim
,
Y. S.
,
Machado
,
T. A.
,
Quirin
,
S.
,
Benson
,
B.
,
Kadmon
,
J.
,
Raja
,
C.
,
Chibukhchyan
,
A.
,
Ramakrishnan
,
C.
,
Inoue
,
M.
,
Shane
,
J. C.
,
McKnight
,
D. J.
,
Yoshizawa
,
S.
,
Kato
,
H. E.
,
Ganguli
,
S.
, &
Deisseroth
,
K.
(
2019
).
Cortical layer–specific critical dynamics triggering perception
.
Science
,
365
(
6453
),
eaaw5202
. ,
[PubMed]
Martin
,
A.
,
Ochagavia
,
M. E.
,
Rabasa
,
L. C.
,
Miranda
,
J.
,
Fernandez-de-Cossio
,
J.
, &
Bringas
,
R.
(
2010
).
BisoGenet: A new tool for gene network building, visualization and analysis
.
BMC Bioinformatics
,
11
,
91
. ,
[PubMed]
Meunier
,
D.
,
Lambiotte
,
R.
,
Fornito
,
A.
,
Ersche
,
K. D.
, &
Bullmore
,
E. T.
(
2009
).
Hierarchical modularity in human brain functional networks
.
Frontiers in Neuroinformatics
,
3
,
37
. ,
[PubMed]
Migault
,
G.
,
van der Plas
,
T. L.
,
Trentesaux
,
H.
,
Panier
,
T.
,
Candelier
,
R.
,
Proville
,
R.
,
Englitz
,
B.
,
Debrégeas
,
G.
, &
Bormuth
,
V.
(
2018
).
Whole-brain calcium imaging during physiological vestibular stimulation in larval zebrafish
.
Current Biology
,
28
(
23
),
3723
3735
. ,
[PubMed]
Miri
,
A.
,
Daie
,
K.
,
Burdine
,
R. D.
,
Aksay
,
E.
, &
Tank
,
D. W.
(
2011
).
Regression-based identification of behavior-encoding neurons during large-scale optical imaging of neural activity at cellular resolution
.
Journal of Neurophysiology
,
105
(
2
),
964
980
. ,
[PubMed]
Mitchell
,
S. M.
,
Lange
,
S.
, &
Brus
,
H.
(
2013
).
Gendered citation patterns in international relations journals
.
International Studies Perspectives
,
14
(
4
),
485
492
.
Miyawaki
,
A.
,
Llopis
,
J.
,
Heim
,
R.
,
McCaffery
,
J. M.
,
Adams
,
J. A.
,
Ikura
,
M.
, &
Tsien
,
R. Y.
(
1997
).
Fluorescent indicators for Ca2+ based on green fluorescent proteins and calmodulin
.
Nature
,
388
(
6645
),
882
887
. ,
[PubMed]
Mölter
,
J.
,
Avitan
,
L.
, &
Goodhill
,
G. J.
(
2018
).
Detecting neural assemblies in calcium imaging data
.
BMC Biology
,
16
(
1
),
143
. ,
[PubMed]
Mu
,
Y.
,
Bennett
,
D. V.
,
Rubinov
,
M.
,
Narayan
,
S.
,
Yang
,
C.-T.
,
Tanimoto
,
M.
,
Mensh
,
B. D.
,
Looger
,
L. L.
, &
Ahrens
,
M. B.
(
2019
).
Glia accumulate evidence that actions are futile and suppress unsuccessful behavior
.
Cell
,
178
(
1
),
27
43
. ,
[PubMed]
Muldoon
,
S. F.
, &
Bassett
,
D. S.
(
2016
).
Network and multilayer network approaches to understanding human brain dynamics
.
Philosophy of Science
,
83
(
5
),
710
720
.
Muldoon
,
S. F.
,
Bridgeford
,
E. W.
, &
Bassett
,
D. S.
(
2016
).
Small-world propensity and weighted brain networks
.
Scientific Reports
,
6
,
22057
. ,
[PubMed]
Muller
,
L.
,
Destexhe
,
A.
, &
Rudolph-Lilith
,
M.
(
2014
).
Brain networks: Small-worlds, after all?
New Journal of Physics
,
16
(
10
),
105004
.
Munch
,
E.
(
2017
).
A user’s guide to topological data analysis
.
Journal of Learning Analytics
,
4
(
2
),
47
61
.
Murphy
,
A. C.
,
Gu
,
S.
,
Khambhati
,
A. N.
,
Wymbs
,
N. F.
,
Grafton
,
S. T.
,
Satterthwaite
,
T. D.
, &
Bassett
,
D. S.
(
2016
).
Explicitly linking regional activation and function connectivity: Community structure of weighted networks with continuous annotation
.
arXiv:1611.07962
.
Murugesan
,
G. K.
,
Ganesh
,
C.
,
Nalawade
,
S.
,
Davenport
,
E. M.
,
Wagner
,
B.
,
Kim
,
W. H.
, &
Maldjian
,
J. A.
(
2020
).
BrainNET: Inference of brain network topology using machine learning
.
Brain Connectivity
,
10
(
8
),
422
435
. ,
[PubMed]
Nakai
,
J.
,
Ohkura
,
M.
, &
Imoto
,
K.
(
2001
).
A high signal-to-noise Ca(2+) probe composed of a single green fluorescent protein
.
Nature Biotechnology
,
19
(
2
),
137
141
. ,
[PubMed]
Naumann
,
E. A.
,
Fitzgerald
,
J. E.
,
Dunn
,
T. W.
,
Rihel
,
J.
,
Sompolinsky
,
H.
, &
Engert
,
F.
(
2016
).
From whole-brain data to functional circuit models: The zebrafish optomotor response
.
Cell
,
167
(
4
),
947
960
. ,
[PubMed]
Nelson
,
C. J.
, &
Bonner
,
S.
(
2021
).
Neuronal graphs: A graph theory primer for microscopic, functional networks of neurons recorded by calcium imaging
.
Frontiers in Neural Circuits
,
15
,
662882
. ,
[PubMed]
Nguyen
,
J. P.
,
Shipley
,
F. B.
,
Linder
,
A. N.
,
Plummer
,
G. S.
,
Liu
,
M.
,
Setru
,
S. U.
,
Shaevitz
,
J. W.
, &
Leifer
,
A. M.
(
2016
).
Whole-brain calcium imaging with cellular resolution in freely behaving Caenorhabditis elegans
.
Proceedings of the National Academy of Sciences
,
113
(
8
),
E1074
E1081
. ,
[PubMed]
Nicolau
,
M.
,
Levine
,
A. J.
, &
Carlsson
,
G.
(
2011
).
Topology based data analysis identifies a subgroup of breast cancers with a unique mutational profile and excellent survival
.
Proceedings of the National Academy of Sciences
,
108
(
17
),
7265
7270
. ,
[PubMed]
Nicosia
,
V.
,
Vértes
,
P. E.
,
Schafer
,
W. R.
,
Latora
,
V.
, &
Bullmore
,
E. T.
(
2013
).
Phase transition in the economically modeled growth of a cellular nervous system
.
Proceedings of the National Academy of Sciences
,
110
(
19
),
7880
7885
. ,
[PubMed]
Nielson
,
J. L.
,
Paquette
,
J.
,
Liu
,
A. W.
,
Guandique
,
C. F.
,
Tovar
,
C. A.
,
Inoue
,
T.
,
Irvine
,
K.-A.
,
Gensel
,
J. C.
,
Kloke
,
J.
,
Petrossian
,
T. C.
,
Lum
,
P. Y.
,
Carlsson
,
G. E.
,
Manley
,
G. T.
,
Young
,
W.
,
Beattie
,
M. S.
,
Bresnahan
,
J. C.
, &
Ferguson
,
A. R.
(
2015
).
Topological data analysis for discovery in preclinical spinal cord injury and traumatic brain injury
.
Nature Communications
,
6
,
8581
. ,
[PubMed]
Nienborg
,
H.
, &
Cumming
,
B.
(
2010
).
Correlations between the activity of sensory neurons and behavior: How much do they tell us about a neuron’s causality?
Current Opinion in Neurobiology
,
20
(
3
),
376
381
. ,
[PubMed]
Oldfield
,
C. S.
,
Grossrubatscher
,
I.
,
Chávez
,
M.
,
Hoagland
,
A.
,
Huth
,
A. R.
,
Carroll
,
E. C.
,
Prendergast
,
A.
,
Qu
,
T.
,
Gallant
,
J. L.
,
Wyart
,
C.
, &
Isacoff
,
E. Y.
(
2020
).
Experience, circuit dynamics, and forebrain recruitment in larval zebrafish prey capture
.
eLife
,
9
,
e56619
. ,
[PubMed]
Onnela
,
J.-P.
,
Fenn
,
D. J.
,
Reid
,
S.
,
Porter
,
M. A.
,
Mucha
,
P. J.
,
Fricker
,
M. D.
, &
Jones
,
N. S.
(
2012
).
Taxonomies of networks from community structure
.
Physical Review E
,
86
(
3 Pt. 2
),
036104
. ,
[PubMed]
Paninski
,
L.
,
Pillow
,
J.
, &
Lewi
,
J.
(
2007
).
Statistical models for neural encoding, decoding, and optimal stimulus design
.
Progress in Brain Research
,
165
,
493
507
. ,
[PubMed]
Paulk
,
A. C.
,
Kfir
,
Y.
,
Khanna
,
A. R.
,
Mustroph
,
M. L.
,
Trautmann
,
E. M.
,
Soper
,
D. J.
,
Stavisky
,
S. D.
,
Welkenhuysen
,
M.
,
Dutta
,
B.
,
Shenoy
,
K. V.
,
Hochberg
,
L. R.
,
Richardson
,
R. M.
,
Williams
,
Z. M.
, &
Cash
,
S. S.
(
2022
).
Large-scale neural recordings with single neuron resolution using Neuropixels probes in human cortex
.
Nature Neuroscience
,
25
(
2
),
252
263
. ,
[PubMed]
Pedersen
,
M.
,
Zalesky
,
A.
,
Omidvarnia
,
A.
, &
Jackson
,
G. D.
(
2018
).
Multilayer network switching rate predicts brain performance
.
Proceedings of the National Academy of Sciences
,
115
(
52
),
13376
13381
. ,
[PubMed]
Phan-Luong
,
V.
(
2008
).
A framework for integrating information sources under lattice structure
.
Information Fusion
,
9
(
2
),
278
292
.
Ponce-Alvarez
,
A.
,
Jouary
,
A.
,
Privat
,
M.
,
Deco
,
G.
, &
Sumbre
,
G.
(
2018
).
Whole-brain neuronal activity displays crackling noise dynamics
.
Neuron
,
100
(
6
),
1446
1459
. ,
[PubMed]
Poulsen
,
R. E.
,
Scholz
,
L. A.
,
Constantin
,
L.
,
Favre-Bulle
,
I.
,
Vanwalleghem
,
G. C.
, &
Scott
,
E. K.
(
2021
).
Broad frequency sensitivity and complex neural coding in the larval zebrafish auditory system
.
Current Biology
,
31
(
9
),
1977
1987
. ,
[PubMed]
Privat
,
M.
,
Romano
,
S. A.
,
Pietri
,
T.
,
Jouary
,
A.
,
Boulanger-Weill
,
J.
,
Elbaz
,
N.
,
Duchemin
,
A.
,
Soares
,
D.
, &
Sumbre
,
G.
(
2019
).
Sensorimotor transformations in the zebrafish auditory system
.
Current Biology
,
29
(
23
),
4010
4023
. ,
[PubMed]
Quian Quiroga
,
R.
, &
Panzeri
,
S.
(
2009
).
Extracting information from neuronal populations: Information theory and decoding approaches
.
Nature Reviews Neuroscience
,
10
(
3
),
173
185
. ,
[PubMed]
Reimann
,
M. W.
,
Nolte
,
M.
,
Scolamiero
,
M.
,
Turner
,
K.
,
Perin
,
R.
,
Chindemi
,
G.
,
Dłotko
,
P.
,
Levi
,
R.
,
Hess
,
K.
, &
Markram
,
H.
(
2017
).
Cliques of neurons bound into cavities provide a missing link between structure and function
.
Frontiers in Computational Neuroscience
,
11
,
48
. ,
[PubMed]
Riquelme
,
J. L.
, &
Gjorgjieva
,
J.
(
2021
).
Towards readable code in neuroscience
.
Nature Reviews Neuroscience
,
22
(
5
),
257
258
. ,
[PubMed]
Robinson
,
M.
(
2013
).
Understanding networks and their behaviors using sheaf theory
.
2013 IEEE Global Conference on Signal and Information Processing
.
Robinson
,
M.
(
2017
).
Sheaves are the canonical data structure for sensor integration
.
Information Fusion
,
36
,
208
224
.
Rombach
,
M. P.
,
Porter
,
M. A.
,
Fowler
,
J. H.
, &
Mucha
,
P. J.
(
2014
).
Core-periphery structure in networks
.
SIAM Journal on Applied Mathematics
,
74
(
1
),
167
190
.
Rubin
,
A.
,
Sheintuch
,
L.
,
Brande-Eilat
,
N.
,
Pinchasof
,
O.
,
Rechavi
,
Y.
,
Geva
,
N.
, &
Ziv
,
Y.
(
2019
).
Revealing neural correlates of behavior without behavioral measurements
.
Nature Communications
,
10
(
1
),
4745
. ,
[PubMed]
Saggar
,
M.
,
Sporns
,
O.
,
Gonzalez-Castillo
,
J.
,
Bandettini
,
P. A.
,
Carlsson
,
G.
,
Glover
,
G.
, &
Reiss
,
A. L.
(
2018
).
Towards a new approach to reveal dynamical organization of the brain using topological data analysis
.
Nature Communications
,
9
(
1
),
1399
. ,
[PubMed]
Schroeder
,
M. E.
,
Bassett
,
D. S.
, &
Meaney
,
D. F.
(
2022
).
A multilayer network model of neuron-astrocyte populations in vitro reveals mGluR5 inhibition is protective following traumatic injury
.
Network Neuroscience
,
6
(
2
),
499
527
. ,
[PubMed]
Shainer
,
I.
,
Kuehn
,
E.
,
Laurell
,
E.
,
Kassar
,
M. A.
,
Mokayes
,
N.
,
Sherman
,
S.
,
Larsch
,
J.
,
Kunst
,
M.
, &
Baier
,
H.
(
2022
).
A single-cell resolution gene expression atlas of the larval zebrafish brain
.
bioRxiv
.
Shih
,
C.-T.
,
Lin
,
Y.-J.
,
Wang
,
C.-T.
,
Wang
,
T.-Y.
,
Chen
,
C.-C.
,
Su
,
T.-S.
,
Lo
,
C.-C.
, &
Chiang
,
A.-S.
(
2020
).
Diverse community structures in the neuronal-level connectome of the Drosophila brain
.
Neuroinformatics
,
18
(
2
),
267
281
. ,
[PubMed]
Shih
,
C.-T.
,
Sporns
,
O.
,
Yuan
,
S.-L.
,
Su
,
T.-S.
,
Lin
,
Y.-J.
,
Chuang
,
C.-C.
,
Wang
,
T.-Y.
,
Lo
,
C.-C.
,
Greenspan
,
R. J.
, &
Chiang
,
A.-S.
(
2015
).
Connectomics-based analysis of information flow in the Drosophila brain
.
Current Biology
,
25
(
10
),
1249
1258
. ,
[PubMed]
Singh
,
G.
,
Mémoli
,
F.
, &
Carlsson
,
G. E.
(
2007
).
Topological methods for the analysis of high dimensional data sets and 3D object recognition
. In
Eurographics Symposium on Point-Based Graphics
(pp.
91
100
).
Sizemore
,
A.
,
Giusti
,
C.
, &
Bassett
,
D. S.
(
2016
).
Classification of weighted networks through mesoscale homological features
.
Journal of Complex Networks
,
5
(
2
),
245
273
.
Sizemore
,
A. E.
,
Giusti
,
C.
,
Kahn
,
A.
,
Vettel
,
J. M.
,
Betzel
,
R. F.
, &
Bassett
,
D. S.
(
2018
).
Cliques and cavities in the human connectome
.
Journal of Computational Neuroscience
,
44
(
1
),
115
145
. ,
[PubMed]
Sizemore
,
A. E.
,
Phillips-Cremins
,
J. E.
,
Ghrist
,
R.
, &
Bassett
,
D. S.
(
2019
).
The importance of the whole: Topological data analysis for the network neuroscientist
.
Network Neuroscience
,
3
(
3
),
656
673
. ,
[PubMed]
Sizemore Blevins
,
A.
, &
Bassett
,
D. S.
(
2020
).
Reorderability of node-filtered order complexes
.
Physical Review E
,
101
(
5-1
),
052311
. ,
[PubMed]
Smith
,
P. F.
(
2018
).
On the application of multivariate statistical and data mining analyses to data in neuroscience
.
Journal of Undergraduate Neuroscience Education: JUNE
,
16
(
2
),
R20
R32
.
[PubMed]
Sood
,
G.
, &
Laohaprapanon
,
S.
(
2018
).
Predicting race and ethnicity from the sequence of characters in a name
.
arXiv:1805.02109
.
Spivak
,
D. I.
(
2009
).
Higher-dimensional models of networks
.
arXiv:0909.4314
.
Sporns
,
O.
(
2018
).
Graph theory methods: Applications in brain networks
.
Dialogues in Clinical Neuroscience
,
20
(
2
),
111
121
. ,
[PubMed]
Steinmetz
,
N. A.
,
Zatka-Haas
,
P.
,
Carandini
,
M.
, &
Harris
,
K. D.
(
2019
).
Distributed coding of choice, action and engagement across the mouse brain
.
Nature
,
576
(
7786
),
266
273
. ,
[PubMed]
Stevenson
,
A. J.
,
Vanwalleghem
,
G.
,
Stewart
,
T. A.
,
Condon
,
N. D.
,
Lloyd-Lewis
,
B.
,
Marino
,
N.
,
Putney
,
J. W.
,
Scott
,
E. K.
,
Ewing
,
A. D.
, &
Davis
,
F. M.
(
2020
).
Multiscale imaging of basal cell dynamics in the functionally mature mammary gland
.
Proceedings of the National Academy of Sciences
,
117
(
43
),
26822
26832
. ,
[PubMed]
Stevenson
,
I. H.
, &
Kording
,
K. P.
(
2011
).
How advances in neural recording affect data analysis
.
Nature Neuroscience
,
14
(
2
),
139
142
. ,
[PubMed]
Stobb
,
M.
,
Peterson
,
J. M.
,
Mazzag
,
B.
, &
Gahtan
,
E.
(
2012
).
Graph theoretical model of a sensorimotor connectome in zebrafish
.
PLoS ONE
,
7
(
5
),
e37292
. ,
[PubMed]
Stringer
,
C.
,
Pachitariu
,
M.
,
Steinmetz
,
N.
,
Reddy
,
C. B.
,
Carandini
,
M.
, &
Harris
,
K. D.
(
2019
).
Spontaneous behaviors drive multidimensional, brainwide activity
.
Science
,
364
(
6437
),
255
. ,
[PubMed]
Sugihara
,
G.
,
May
,
R.
,
Ye
,
H.
,
Hsieh
,
C.-H.
,
Deyle
,
E.
,
Fogarty
,
M.
, &
Munch
,
S.
(
2012
).
Detecting causality in complex ecosystems
.
Science
,
338
(
6106
),
496
500
. ,
[PubMed]
Tainton-Heap
,
L. A. L.
,
Kirszenblat
,
L. C.
,
Notaras
,
E. T.
,
Grabowska
,
M. J.
,
Jeans
,
R.
,
Feng
,
K.
,
Shaw
,
P. J.
, &
van Swinderen
,
B.
(
2021
).
A paradoxical kind of sleep in Drosophila melanogaster
.
Current Biology
,
31
(
3
),
578
590
. ,
[PubMed]
Tajima
,
S.
,
Mita
,
T.
,
Bakkum
,
D. J.
,
Takahashi
,
H.
, &
Toyoizumi
,
T.
(
2017
).
Locally embedded presages of global network bursts
.
Proceedings of the National Academy of Sciences
,
114
(
36
),
9517
9522
. ,
[PubMed]
Tajima
,
S.
,
Yanagawa
,
T.
,
Fujii
,
N.
, &
Toyoizumi
,
T.
(
2015
).
Untangling brain-wide dynamics in consciousness by cross-embedding
.
PLoS Computational Biology
,
11
(
11
),
e1004537
. ,
[PubMed]
Tallon-Baudry
,
C.
,
Bertrand
,
O.
, &
Fischer
,
C.
(
2001
).
Oscillatory synchrony between human extrastriate areas during visual short-term memory maintenance
.
Journal of Neuroscience
,
21
(
20
),
RC177
. ,
[PubMed]
Tang
,
E.
,
Giusti
,
C.
,
Baum
,
G. L.
,
Gu
,
S.
,
Pollock
,
E.
,
Kahn
,
A. E.
,
Roalf
,
D. R.
,
Moore
,
T. M.
,
Ruparel
,
K.
,
Gur
,
R. C.
,
Gur
,
R. E.
,
Satterthwaite
,
T. D.
, &
Bassett
,
D. S.
(
2017
).
Developmental increases in white matter network controllability support a growing diversity of brain dynamics
.
Nature Communications
,
8
(
1
),
1252
. ,
[PubMed]
Testa-Silva
,
G.
,
Loebel
,
A.
,
Giugliano
,
M.
,
de Kock
,
C. P. J.
,
Mansvelder
,
H. D.
, &
Meredith
,
R. M.
(
2012
).
Hyperconnectivity and slow synapses during early development of medial prefrontal cortex in a mouse model for mental retardation and autism
.
Cerebral Cortex
,
22
(
6
),
1333
1342
. ,
[PubMed]
Theunissen
,
F.
, &
Miller
,
J. P.
(
1995
).
Temporal encoding in nervous systems: A rigorous definition
.
Journal of Computational Neuroscience
,
2
(
2
),
149
162
. ,
[PubMed]
Thompson
,
A. W.
,
Vanwalleghem
,
G. C.
,
Heap
,
L. A.
, &
Scott
,
E. K.
(
2016
).
Functional profiles of visual-, auditory-, and water flow-responsive neurons in the zebrafish tectum
.
Current Biology
,
26
(
6
),
743
754
. ,
[PubMed]
Thompson
,
W. H.
,
Brantefors
,
P.
, &
Fransson
,
P.
(
2017
).
From static to temporal network theory: Applications to functional brain connectivity
.
Network Neuroscience
,
1
(
2
),
69
99
. ,
[PubMed]
Tompson
,
S. H.
,
Kahn
,
A. E.
,
Falk
,
E. B.
,
Vettel
,
J. M.
, &
Bassett
,
D. S.
(
2020
).
Functional brain network architecture supporting the learning of social networks in humans
.
NeuroImage
,
210
,
116498
. ,
[PubMed]
Torres
,
L.
,
Blevins
,
A. S.
,
Bassett
,
D. S.
, &
Eliassi-Rad
,
T.
(
2020
).
The why, how, and when of representations for complex systems
.
arXiv:2006.02870
.
Trinh
,
M. H.
,
Van Nguyen
,
C.
,
Lim
,
Y.-H.
, &
Ahn
,
H.-S.
(
2018
).
Matrix-weighted consensus and its applications
.
Automatica
,
89
,
415
419
.
Triplett
,
M. A.
,
Pujic
,
Z.
,
Sun
,
B.
,
Avitan
,
L.
, &
Goodhill
,
G. J.
(
2020
).
Model-based decoupling of evoked and spontaneous neural activity in calcium imaging data
.
PLoS Computational Biology
,
16
(
11
),
e1008330
. ,
[PubMed]
Tsien
,
R. Y.
,
Pozzan
,
T.
, &
Rink
,
T. J.
(
1982
).
Calcium homeostasis in intact lymphocytes: Cytoplasmic free calcium monitored with a new, intracellularly trapped fluorescent indicator
.
Journal of Cell Biology
,
94
(
2
),
325
334
. ,
[PubMed]
Vaiana
,
M.
, &
Muldoon
,
S. F.
(
2020
).
Multilayer brain networks
.
Journal of Nonlinear Science
,
30
(
5
),
2147
2169
.
van Veen
,
H.
,
Saul
,
N.
,
Eargle
,
D.
, &
Mangham
,
S.
(
2019
).
Kepler Mapper: A flexible Python implementation of the Mapper algorithm
.
Journal of Open Source Software
,
4
(
42
),
1315
.
Vanwalleghem
,
G. C.
,
Ahrens
,
M. B.
, &
Scott
,
E. K.
(
2018
).
Integrative whole-brain neuroscience in larval zebrafish
.
Current Opinion in Neurobiology
,
50
,
136
145
. ,
[PubMed]
Vanwalleghem
,
G. C.
,
Heap
,
L. A.
, &
Scott
,
E. K.
(
2017
).
A profile of auditory-responsive neurons in the larval zebrafish brain
.
Journal of Comparative Neurology
,
525
(
14
),
3031
3043
. ,
[PubMed]
Vanwalleghem
,
G. C.
,
Schuster
,
K.
,
Taylor
,
M. A.
,
Favre-Bulle
,
I. A.
, &
Scott
,
E. K.
(
2020
).
Brain-wide mapping of water flow perception in zebrafish
.
Journal of Neuroscience
,
40
(
21
),
4130
4144
. ,
[PubMed]
Vértes
,
P. E.
,
Alexander-Bloch
,
A.
, &
Bullmore
,
E. T.
(
2014
).
Generative models of rich clubs in Hebbian neuronal networks and large-scale human brain networks
.
Philosophical Transactions of the Royal Society of London B: Biological Sciences
,
369
(
1653
),
20130531
. ,
[PubMed]
Vértes
,
P. E.
,
Alexander-Bloch
,
A. F.
,
Gogtay
,
N.
,
Giedd
,
J. N.
,
Rapoport
,
J. L.
, &
Bullmore
,
E. T.
(
2012
).
Simple models of human brain functional networks
.
Proceedings of the National Academy of Sciences
,
109
(
15
),
5868
5873
. ,
[PubMed]
Vishwanathan
,
A.
,
Ramirez
,
A. D.
,
Wu
,
J.
,
Sood
,
A.
,
Yang
,
R.
,
Kemnitz
,
N.
,
Ih
,
D.
,
Turner
,
N.
,
Lee
,
K.
,
Tartavull
,
I.
,
Silversmith
,
W. M.
,
Jordan
,
C. S.
,
David
,
C.
,
Bland
,
D.
,
Goldman
,
M. S.
,
Aksay
,
E. R. F.
,
Sebastian Seung
,
H.
, &
the EyeWirers
. (
2022
).
Predicting modular functions and neural coding of behavior from a synaptic wiring diagram
.
bioRxiv
.
Vladimirov
,
N.
,
Wang
,
C.
,
Höckendorf
,
B.
,
Pujala
,
A.
,
Tanimoto
,
M.
,
Mu
,
Y.
,
Yang
,
C.-T.
,
Wittenbach
,
J. D.
,
Freeman
,
J.
,
Preibisch
,
S.
,
Koyama
,
M.
,
Keller
,
P. J.
, &
Ahrens
,
M. B.
(
2018
).
Brain-wide circuit interrogation at the cellular level guided by online analysis of neuronal function
.
Nature Methods
,
15
(
12
),
1117
1125
. ,
[PubMed]
Warp
,
E.
,
Agarwal
,
G.
,
Wyart
,
C.
,
Friedmann
,
D.
,
Oldfield
,
C. S.
,
Conner
,
A.
,
Del Bene
,
F.
,
Arrenberg
,
A. B.
,
Baier
,
H.
, &
Isacoff
,
E. Y.
(
2012
).
Emergence of patterned activity in the developing zebrafish spinal cord
.
Current Biology
,
22
(
2
),
93
102
. ,
[PubMed]
Watanakeesuntorn
,
W.
,
Takahashi
,
K.
,
Ichikawa
,
K.
,
Park
,
J.
,
Sugihara
,
G.
,
Takano
,
R.
,
Haga
,
J.
, &
Pao
,
G. M.
(
2020
).
Massively parallel causal inference of whole brain dynamics at single neuron resolution
.
2020 IEEE 26th International Conference on Parallel and Distributed Systems (ICPADS)
,
196
205
.
Wiener
,
N.
(
1956
).
The theory of prediction: Modern mathematics for engineers
.
Xu
,
J.
,
Wickramarathne
,
T. L.
, &
Chawla
,
N. V.
(
2016
).
Representing higher-order dependencies in networks
.
Science Advances
,
2
(
5
),
e1600028
. ,
[PubMed]
Yang
,
J.
,
McAuley
,
J.
, &
Leskovec
,
J.
(
2013
).
Community detection in networks with node attributes
.
2013 IEEE 13th International Conference on Data Mining
.
Yoo
,
J.
,
Kim
,
E. Y.
,
Ahn
,
Y. M.
, &
Ye
,
J. C.
(
2016
).
Topological persistence vineyard for dynamic functional brain connectivity during resting and gaming stages
.
Journal of Neuroscience Methods
,
267
,
1
13
. ,
[PubMed]
Yoon
,
H. R.
, &
Ghrist
,
R.
(
2020
).
Persistence by parts: Multiscale feature detection via distributed persistent homology
.
arXiv:2001.01623
.
Yu
,
M.
,
Engels
,
M. M. A.
,
Hillebrand
,
A.
,
van Straaten
,
E. C. W.
,
Gouw
,
A. A.
,
Teunissen
,
C.
,
van der Flier
,
W. M.
,
Scheltens
,
P.
, &
Stam
,
C. J.
(
2017
).
Selective impairment of hippocampus and posterior hub areas in Alzheimer’s disease: An MEG-based multiplex network study
.
Brain
,
140
(
5
),
1466
1485
. ,
[PubMed]
Zhang
,
D. D.
,
Lee
,
H. F.
,
Wang
,
C.
,
Li
,
B.
,
Pei
,
Q.
,
Zhang
,
J.
, &
An
,
Y.
(
2011
).
The causality analysis of climate change and large-scale human crisis
.
Proceedings of the National Academy of Sciences
,
108
(
42
),
17296
17301
. ,
[PubMed]
Zhang
,
X.
,
Braun
,
U.
,
Harneit
,
A.
,
Zang
,
Z.
,
Geiger
,
L. S.
,
Betzel
,
R. F.
,
Chen
,
J.
,
Schweiger
,
J. I.
,
Schwarz
,
K.
,
Reinwald
,
J. R.
,
Fritze
,
S.
,
Witt
,
S.
,
Rietschel
,
M.
,
Nöthen
,
M. M.
,
Degenhardt
,
F.
,
Schwarz
,
E.
,
Hirjak
,
D.
,
Meyer-Lindenberg
,
A.
,
Bassett
,
D. S.
, &
Tost
,
H.
(
2021
).
Generative network models of altered structural brain connectivity in schizophrenia
.
NeuroImage
,
225
,
117510
. ,
[PubMed]
Zheng
,
Z.
,
Lauritzen
,
J. S.
,
Perlman
,
E.
,
Robinson
,
C. G.
,
Nichols
,
M.
,
Milkie
,
D.
,
Torrens
,
O.
,
Price
,
J.
,
Fisher
,
C. B.
,
Sharifi
,
N.
,
Calle-Schuler
,
S. A.
,
Kmecova
,
L.
,
Ali
,
I. J.
,
Karsh
,
B.
,
Trautman
,
E. T.
,
Bogovic
,
J. A.
,
Hanslovsky
,
P.
,
Jefferis
,
G. S. X. E.
,
Kazhdan
,
M.
, …
Bock
,
D. D.
(
2018
).
A complete electron microscopy volume of the brain of adult Drosophila melanogaster
.
Cell
,
174
(
3
),
730
743
. ,
[PubMed]
Zhou
,
D.
,
Cornblath
,
E. J.
,
Stiso
,
J.
,
Teich
,
E. G.
,
Dworkin
,
J. D.
,
Blevins
,
A. S.
, &
Bassett
,
D. S.
(
2020
).
Gender diversity statement and code notebook v1.0
.
Zenodo
.

Author notes

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

Handling Editor: Alex Fornito

This is an open-access article distributed under the terms of the Creative Commons Attribution 4.0 International License, which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properly cited. For a full description of the license, please visit https://creativecommons.org/licenses/by/4.0/legalcode.